You are on page 1of 83

Advances in Colloid and Interface Science 81 1999.

167 249

Contact angle measurement and contact angle interpretation


D.Y. Kwok1, A.W. NeumannU
Department of Mechanical and Industrial Engineering, Uni ersity of Toronto, 5 Kings College Road, Toronto, Ontario, Canada M5S 3G8

Abstract Recent progress in the correlation of contact angles with solid surface tensions are summarized. The measurements of meaningful contact angles in terms of surface energetics are also discussed. It is shown that the controversy with respect to measurement and interpretation of contact angles are due to the fact that some or all. of the assumptions made in all energetic approaches are violated when contact angles are measured and processed. For a large number of polar and non-polar liquids on different solid surfaces, the liquid vapor surface tension times cosine of the contact angle, l cos , is shown to depend only on the liquid vapor surface tension l , and the solid vapor surface tension s when the appropriate experimental techniques and procedures are used. Equations which follow these experimental patterns and which allow the determination of solid surface tensions from contact angles are discussed. Universality of these experimental contact angle patterns is illustrated; other reasons which may cause data to deviate from the patterns slightly are discussed. It is found that surface tension component approaches do not reflect physical reality. Assuming the fact that solid surface tension is constant for one and the same solid surface, experimental contact angle patterns are employed to deduce a functional relationship to be used in conjunction with Youngs equation for determining solid surface tensions. The explicit form of such a relation is obtained by modifying Berthelots rule together with experimental data; essentially constant solid surface tension values are obtained, independent of liquid surface tension and molecular structure. A new combining rule is also derived based on an expression similar to one used in molecular theory; such a combining rule
U

Corresponding author. Tel.: q1-416-978-1270; fax: q1-416-978-7753. E-mail address: neumann@me.utoronto.ca A.W. Neumann. 1 This paper represents, in part, the Ph.D. thesis of D.Y. Kwok. Present address: Department of Chemical Engineering Massachusetts Institute of Technology, Cambridge, MA 02139, USA. Tel.: q1-617-253-6482; fax: q1-617-258-5042; e-mail: dykwok@mit.edu

0001-8686r99r$ - see front matter 1999 Elsevier Science B.V. All rights reserved. PII: S 0 0 0 1 - 8 6 8 6 9 9 . 0 0 0 8 7 - 6

168

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

should allow a better understanding of the molecular interactions between unlike solid liquid pairs from like pairs. Existing static contact angles for 34 different types of solid surfaces from Zisman et al. are evaluated in terms of their solid surface tensions using experimental contact angle patterns. A FORTRAN computer program has been implemented to automate these procedures. It is found that literature contact angles do not have to be discarded completely; they can be used to determine solid surface tensions, with caution. The surface tensions for the 34 solid surfaces from Zisman et al. are also reported. 1999 Elsevier Science B.V. All rights reserved.
Keywords: Contact angle; Solid surface tension; Surface tension component; Combining rule; Drop shape method

Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169 2. Contact angle measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172 2.1. Experimental contact angle patterns . . . . . . . . . . . . . . . . . . . . . . . . . . . 172 2.2. Experimental contant angle patterns from a goniometer . . . . . . . . . . . . . . 175 2.3. Low-rate dynamic advancing. contact angles by ADSA-P . . . . . . . . . . . . . 176 2.3.1. Experimental procedures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176 2.3.2. Inert non-polar. surfaces: FC-722-coated mica surface . . . . . . . . . . . 179 2.3.3. Non-inert polar. surfaces: polypropene-alt-N- n-hexyl.maleimide. and polypropene-alt-N- n-propyl.maleimide. . . . . . . . . . . . . . . . . . . . . 182 2.3.4. Other non-polar and polar surfaces . . . . . . . . . . . . . . . . . . . . . . . . 189 2.4. Universality of contact angle patterns . . . . . . . . . . . . . . . . . . . . . . . . . . 191 2.4.1. Reasons of deviation from smoothness . . . . . . . . . . . . . . . . . . . . . . 194 2.5. Criteria for calculations of surface energetics . . . . . . . . . . . . . . . . . . . . . 196 2.5.1. Accepted assumptions for calculations of surface energetics . . . . . . . . 196 2.5.2 Experimental criteria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198 3. Contact angle interpretation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200 3.1. Surface tension component approaches . . . . . . . . . . . . . . . . . . . . . . . . . 202 3.1.1. Fowkes approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202 3.1.2. Owens-Wendt-Kaelble approach . . . . . . . . . . . . . . . . . . . . . . . . . 203 3.1.3. Lifshitz-van der Waals r acid-base van Oss. approach . . . . . . . . . . . 204 3.2. Equation of state approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205 3.2.1. Antonows rule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205 3.2.2. Berthelots geometric mean. combining rule . . . . . . . . . . . . . . . . . 206 3.2.3. Equation of state approach: modified Berthelots rule . . . . . . . . . . . . 209 3.2.4. Equation of state approach: alternative formulation . . . . . . . . . . . . . 214 3.2.5. Comparison with the original equation of state formulation . . . . . . . . 218 3.2.6. Predicitive power of the equation of state approach . . . . . . . . . . . . . 221 4. Evaluation of existing contact angle data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222 5. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235 6. Perspective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236 7. Future development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

169

Appendix A Appendix B Appendix C Appendix D References .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. 238 . 239 . 240 . 242 . 246

1. Introduction The determination of solid vapor s . and solid liquid sl . interfacial tensions is of importance in a wide range of problems in pure and applied science. Because of the difficulties involved in measuring directly the surface tension involving a solid phase, indirect approaches are called for: Several independent approaches have been used to estimate solid surface tensions, including direct force measurements w1 9x; contact angles w10 17x; capillary penetration into columns of particle powder w18 21x; sedimentation of particles w22 25x; solidification front interaction with particles w26 33x; film flotation w34 38x; gradient theory w39 42x; Lifshitz theory of van der Waals forces w42 45x; and theory of molecular interactions w46 49x. Among these methods, contact angle measurements are believed to be the simplest. Contact angle measurement is easily performed by establishing the tangent angle. of a liquid drop with a solid surface at the base. The attractiveness of using contact angles to estimate the solid vapor and solid liquid interfacial tensions is due to the relative ease with which contact angles can be measured on suitably prepared solid surfaces. It will become apparent later that this seeming simplicity is, however, very misleading. The possibility of estimating solid surface tensions from contact angles relies on a relation which has been recognized by Young w50x in 1805. The contact angle of a liquid drop on a solid surface is defined by the mechanical equilibrium of the drop under the action of three interfacial tensions Fig. 1.: solid vapor, s , solid liquid, sl , and liquid vapor, l . This equilibrium relation is known as Youngs equation:
l cos Y

sl

1.

where

is the Young contact angle, i.e. a contact angle which can be inserted

Fig. 1. Schematic of a sessile-drop contact angle system.

170

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

into Youngs equation. It will become apparent later that the experimentally accessible contact angles may or may not be equal to Y . Youngs Eq. 1. contains only two measurable quantities, the contact angle and the liquid vapor surface tension, l . In order to determine s and sl , an additional relation relating these quantities must be sought. Nevertheless, Eq. 1. suggests that the observation of the equilibrium contact angles of liquids on solids may be a starting point for investigating the solid surface tensions, s and sl . This has inspired many studies which attempt to develop methodologies for determining solid surface tensions. A common feature of these approaches is the assumption that contact angle measurement is a trivial task. Since l , s and sl are thermodynamic properties of the liquid and solid, Eq. 1. implies a single, unique contact angle; in practice, however, contact angle phenomena are complicated w51 53x. In particular, the contact angle made by an advancing liquid a . and that made by a receding liquid r . are not identical; nearly all solid surfaces exhibit contact angle hysteresis, H the difference between . a and r : Hs
a

2.

Contact angle hysteresis can be due to roughness and heterogeneity of a solid surface. If roughness is the primary cause, then the measured contact angles are meaningless in terms of Youngs equation. On very rough surfaces, contact angles are larger than on chemically identical, smooth surfaces w21x. Obviously, interpreting such angles in terms of Eq. 1. would lead to erroneous results because the contact angle would inevitably reflect surface topography, rather than exclusively surface energetics. Because of these various complexities, models have been employed to gain a deeper understanding of the thermodynamic status of contact angles. In general, it has been found that the experimentally observed apparent contact angle, , may or may not be equal to the Young contact angle, Y w51,52x: 1. On ideal solid surfaces, there is no contact angle hysteresis and the experimentally observed contact angle is equal to Y . 2. On smooth, but chemically heterogeneous solid surfaces, is not necessarily equal to the thermodynamic equilibrium angle. Nevertheless, the experimental advancing contact angle, a , can be expected to be a good approximation of Y . This has been illustrated using a model of a vertical surface consisting of heterogeneous and smooth strips w51,52x. Therefore, care must be exercised to ensure that the experimental apparent contact angle, , is the advancing contact angle in order to be inserted into the Young equation. While the receding angle on a heterogeneous and smooth surface can also be a Young angle, it is usually found to be non-reproducible often because of sorption of the liquid into the solid and swelling of the solid by the liquid w54x. 3. On rough solid surfaces, no such equality between a and Y exists: all contact angles on rough surfaces are meaningless in terms of Youngs equation.

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

171

The thermodynamic equilibrium angles on rough and heterogeneous surfaces are the so-called Wenzel w55x and Cassie w56 58x angles, respectively. They are not equal to Y w51,52x; furthermore, they are not experimentally accessible quantities. There are as yet no general criteria to answer the question of how smooth a solid surface has to be for surface roughness not to have an effect on the contact angle. This and similar problems can be linked to line tension, which has its own complexities w59x. It is, therefore, of utmost importance to prepare solid surfaces as smooth as possible so that the experimental advancing angles can be a good approximation of Y . In addition to these complexities, penetration of the liquid into the solid, swelling of the solid by the liquid, and chemical reactions can all play a role. For example, swelling of a solid by a liquid w54x can change the chemistry of the solid in an unknown manner and hence affect the values of s , andror sl . Therefore, it is also important to ensure that the solid surfaces are as inert as possible, in order to minimize such effects, by appropriate choice of the liquid. Several contact angle approaches w10 17x, of current interest, were largely inspired by the idea of using Youngs equation for the determination of surface energetics. While these approaches are, logically and conceptually, mutually exclusive, they share, nevertheless, some basic assumptions: 1. All approaches rely on the validity and applicability of Youngs equation for surface energetics from experimental contact angles. 2. Pure liquids are always used; surfactant solutions or mixtures of liquids should not be used, since they would introduce complications due to preferential adsorption. 3. The values of l , s and sl . are assumed to be constant during the experiment, i.e. there should be no physical or chemical reaction between the solid and the liquid. 4. The liquid surface tensions of the testing liquids should be higher than the anticipated solid surface tension. 5. The values of s going from liquid to liquid are also assumed to be constant, i.e. independent of the liquids used. With respect to the first assumption, one requires the solid surfaces to be rigid, smooth and homogeneous so that Youngs equation is the appropriate equilibrium condition; the experimentally observed contact angles should also be advancing contact angles. However, many attempts have been made in the literature to interpret surface energetics of solids, which are not rigid e.g. gels w60x. and not smooth e.g. biological surfaces w61x., in conjunction with Youngs equation. Clearly, these results are open to question, since Youngs equation may not be valid or applicable in these situations. With respect to the other assumptions, the solid surfaces should be as inert as possible so that effects, such as swelling and chemical reactions are minimized. In order to assure that the experimentally measured contact angles do not violate any of the above assumptions, one requires careful experimentation and suitable methodology. However, contact angles are typically measured simply by depositing a drop of liquid on a given solid surface, and manually placing a tangent

172

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

to the drop at its base using a so-called goniometer sessile drop technique w62,63x. Apart from the subjectivity of the technique, it normally yields contact angle accuracy of no better than "2 . More important, the technique cannot be expected to reflect the complexities of solid liquid interactions. It will become apparent in this paper that much of the controversy with respect to the interpretation of contact angles in terms of surface energetics lies in the fact that not enough attention is given to the readily accepted assumptions. In this paper, the correlation of contact angles with solid surface tensions will be discussed. Section 2.1 illustrates the experimental contact angle patterns on inert solid surfaces using an automated axisymmetric drop shape analysis; these results will be contrasted with the patterns obtained on non-inert solid surfaces using a conventional goniometer sessile drop technique in Section 2.2. By using the automated axisymmetric drop shape analysis profile ADSA-P. and the appropriate experimental procedures described in Section 2.3, the discrepancy between the two contact angle-patterns is shown to come from the fact that some or all. of the widely accepted assumptions are violated, because the goniometer procedure does not have a built-in quality control mechanism. Section 2.3 illustrates the low-rate dynamic contact angle results for various solid surfaces, as obtained by ADSA-P. The universality of the contact angle patterns will be discussed in Section 2.4. and 2.5 presents a detailed discussion of the criteria for calculations of surface energetics. Equations which follow these experimental patterns and which allow the determination of solid surface tensions from contact angles will be sought in Section 3. Applications of the equation developed in Section 3 will be discussed in Section 4; contact angle data from Zisman et al. w10x on several solid surfaces will be evaluated for energetic calculations. A summary, perspective, and future development will be given, respectively, in Sections 5 7.

2. Contact angle measurement 2.1. Experimental contact angle patterns On carefully prepared solid surfaces, Li et al. w64,65x have performed static advancing. contact angle experiments using an automated axisymmetric drop shape analysis profile ADSA-P.. ADSA-P w66x is a technique to determine liquid fluid interfacial tensions and contact angles from the shape of axisymmetric menisci, i.e. from sessile as well as pendant drops. Assuming that the experimental drop is Laplacian and axisymmetric, ADSA-P finds the theoretical profile that best matches the drop profile extracted from the image of a real drop, from which the surface tension, contact angle, drop volume, surface area and three-phase contact radius can be computed. The strategy employed is to fit the shape of an experimental drop to a theoretical drop profile according to the Laplace equation of capillarity, using surfacerinterfacial tension as an adjustable parameter. The best fit identifies the correct surfacerinterfacial tension from which the contact angle can be determined by a numerical integration of the Laplace equation. Details can

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

173

be found elsewhere w67 69x. It has been found w64,65x that a contact angle accuracy of better than "0.3 can be obtained on well-prepared solid surfaces. In the experiments of Li et al. w64,65x, static advancing. contact angles were measured by supplying test liquids from below the surface into the sessile drop, using a motor-driven syringe device. A hole of approximately 1 mm in the centre of each solid surface was required to facilitate such procedures. Liquid was pumped slowly into the drop from below the surface until the three-phase contact radius was approximately 0.4 cm. After the motor was stopped, the sessile drop was allowed to relax for approximately 30 s to reach equilibrium. Then three pictures of this sessile drop were taken successively at intervals of 30 s. More liquid was then pumped into the drop until it reached another desired size, and the above procedure was repeated w64x. These procedures ensure that the measured static contact angles are indeed the advancing contact angles. Three carefully prepared solid surfaces were used in Li et al.s study: they are FC-721-coated mica, Teflon FEP fluorinated ethylene propylene. heat-pressed against quartz glass slides and polyethylene terephthalate . PET. FC-721 is a 3M Fluorad brand antimigration coating designed to prevent the creep of lubricating oils out of bearings. The FC-721-coated mica was prepared by a dip-coating technique. Teflon FEP fluorinated ethylene propylene. surfaces were prepared by a heat-pressing method. The material was cut to 2 = 4 cm, sandwiched between two glass slides, and heat-pressed by a jig in an oven. Polyethylene terephthalate . PET is the condensation product of ethylene glycol and terephthalic acid. The surface of PET films were exceedingly smooth as received and were cleaned before measurement. Details of the solid surface preparation can be found elsewhere w64x. Fig. 2 shows these contact angle results in a plot of l cos vs. l , for a large number of pure liquids with different molecular properties. It can be seen that, for a given solid surface, l cos changes smoothly and systematically with l . Since the solid surface tension, s , of a given solid surface is expected to be constant, i.e. independent of the choice of the testing liquid used, Fig. 2 implies that l cos depends only on l at constant s . Replacing the solid surface from the hydrophobic FC-721 surface to the hydrophilic PET surface shifts the curve in a very regular manner. Thus, one has to conclude that the values of l cos depend only on l and s , independent of any specific intermolecular forces of the liquids and solids w64,65,70x:
l cos

s f

3.

where f is an as yet unknown function. Because of Youngs equation, the experimental contact angle patterns in Fig. 2 imply, in light of Eq. 3., that the solid liquid interfacial tension sl depends only on the liquid vapor l and solid vapor s surface tensions: combining Youngs equation with Eq. 3. yields
s

sl

s f

4.

174

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Fig. 2. l cos vs. l for three well-prepared inert solid surfaces: FC-721-coated mica, heat-pressed Teflon FEP fluorinated ethylene propylene., and polyethylene terephthalate . PET w64,65x. The smoothness suggests that l cos depends only on l and s .

and hence
sl

y f

. s F

5.

where F is as yet another unknown function. This is indeed in agreement with thermodynamics w71x and the thermodynamic phase rule for capillary systems w72 75x which states that there are only two degrees of freedom for such solid liquid systems. Thus, one can simply change the contact angle by changing either l or s . While the specific intermolecular forces determine the primary surface tensions of liquids and solids, they do not have any additional and independent effects on the contact angles, in the context of Youngs equation. In principle, a plot of cos vs. l , rather than l cos vs. l , can also be used to deduce the functional dependence of Eq. 5.. Historically, a plot of l cos vs. l has been used in conjunction with Youngs equation to argue the fact that as l cos increases and approaches s , sl approaches zero w13,14x. It is to be noted that the experimental contact angle patterns shown in Fig. 2 do not always appear in the literature: curves far less smooth or no unique curves at all are very frequently reported see later.. Such patterns can have a variety of causes. Accurate contact angle measurements require extreme experimental care. Even very minor vibrations can cause advancing contact angles to decrease, resulting in errors of several degrees. Surface roughness can affect the contact angles and make Youngs equation inapplicable. Swelling of a solid by a liquid w54x can change the chemistry of the solid and hence the values of s and in an unpredictable manner. Non-constancy of l , s and sl during the experiment, and non-constancy of s from liquid to liquid can produce scatter in plots of the type of Fig. 2 easily.

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

175

It will be shown below that the present controversy with respect to the experimental contact angle patterns arises from the fact that these patterns are often complex, and cannot be unraveled by the simple goniometer techniques: an example of the experimental patterns, which are very different from those of Fig. 2 will be illustrated, simply by measuring advancing contact angles by a goniometer technique. 2.2. Experimental contact angle patterns from a goniometer In order to compare the experimental contact angle patterns obtained by a goniometer with those by an automated ADSA-P technique shown in Fig. 2, two well-defined copolymers were selected. These copolymers were selected purposely to be not as inert as the FC-721-coated mica, Teflon FEP, and PET surfaces used by Li et al. w64,65x. Two copolymers, polypropene-alt-N- n-propyl.maleimide. and polypropene-altN- n-hexyl.maleimide., i.e. copolymers with different side chains w76x, were used for contact angle measurements. For each copolymer, a 2% solution was prepared using tetrahydrofuran as the solvent. Silicon wafers 100: were selected as the substrate for the copolymer coating. The silicon wafer surfaces were obtained as circular discs of approximately 10 cm diameter and were cut into rectangular shapes of approximately 2 = 3 cm2 . Each rectangular silicon wafer surface was first cleaned with ethanol, acetone, and then soaked in chromic acid for at least 24 h. The cleaned wafer surfaces were rinsed with doubly-distilled water, and dried under a heat lamp before the copolymer coating. A few drops of the 2% copolymerrtetrahydrofuran solution were deposited on the dried silicon wafers inside petri glass dishes overnight; the solution spread and a thin layer of the copolymer formed on the wafer surface after tetrahydrofuran evaporated. This preparation produced good quality coated surfaces, as manifested by light fringes, due to refraction at these surfaces, suggesting that roughness is in the order of nanometers or less. It should be noted that if preparation of the solid surfaces and liquids for contact angle measurements is less careful, impurities can easily contaminate the testing liquids and solids. This would inevitably result in the contact angle measurements of contaminated liquids and solids, rather than the presumed pure liquids and solids. Thirteen liquids were chosen in this study. Selection of these liquids was based on the following criteria: 1. liquids should include a wide range of intermolecular forces; 2. liquids should be non-toxic; and 3. the liquid surface tension should be higher than the anticipated solid surface tension w10,13,21x. They are, in the order of increasing surface tension, cis-decalin, 2,5-dichlorotoluene, ethyl cinnamate, dibenzylamine, dimethyl sulfoxide DMSO., 1-bromonaphthalene, diethylene glycol, ethylene glycol, diiodomethane, 2,2 -thiodiethanol, formamide, glycerol and water. The procedures to measure the advancing contact angles using a goniometer sessile drop technique are the same as those typically used in the literature: a sessile drop of approximately 0.4 0.5 cm radius was formed from

176

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

above through a Teflon capillary. The three-phase contact line of the drop was then slowly advanced by supplying more liquid from above through the capillary which was always kept in contact with the drop. The maximum advancing. contact angles were measured carefully from the left and right side of the drop and subsequently averaged. The above procedures were repeated for five drops on five new surfaces. All readings were then averaged to give an averaged contact angle. Fig. 3 shows the contact angle results for the two copolymers from the goniometer sessile drop measurements in a plot of l cos vs. l . Contrary to the contact angle patterns shown in Fig. 2, considerable scatter is apparent. On a given solid, say the polypropene-alt-N- n-hexyl.maleimide. copolymer, s , is expected to be constant; since the values of l cos here do not appear to give a smooth and systematic change with l , one might argue that l cos cannot be a simple function g of only l and s , but has to depend also on the various specific intermolecular forces of the liquids and solids:
l cos

s g

, dipole dipole, hydrogen bonding, etc..

6.

Thus, because of Youngs equation, this scatter would seem to favor the stipulation of the surface tension component approaches w11,15 17,77 79x that sl depends not only on l and s , but also on the specific intermolecular forces:
s

sl

s g

, dipole dipole, hydrogen bonding, etc..

7.

and hence
sl

y g
l

l s

, dipole dipole, hydrogen bonding, etc.. 8.

s G

, dipole dipole, hydrogen bonding, etc..

where G is an unknown function. It is to be noted that while the contact angle patterns in Fig. 3 can be easily found in the literature, they do not really support the above stipulation of the surface tension component approaches w11,15 17,77 79x. In the next section, it will be shown that this scatter indeed comes from the fact that many of the experimental contact angles in Fig. 3 have violated some or all. assumptions usually made and summarized in the Introduction. Thus, the apparent additional degrees of freedom inferred from the scatter . do not come from the putative independent effect of intermolecular forces on the contact angles: l cos can be shown to change smoothly and systematically with l if suitable experimental techniques and procedures are employed, such as those described below, to delete measurements which violate any of the above assumptions. 2.3. Low-rate dynamic (ad ancing) contact angles by ADSA-P 2.3.1. Experimental procedures Sessile drop contact angle measurements using ADSA-P were performed dynamically, by using a motor-driven syringe to pump liquid steadily into the sessile

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

177

Fig. 3. l cos vs. l for polypropene-alt-N- n-propyl.maleimide. and polypropene-alt-N- nhexyl.maleimide. copolymers. These angles are advancing contact angles measured by a conventional goniometer technique w76x. Due to the scatter, one might argue that l cos cannot be a simple function of l and s .

drop from below the surface w80x. A quartz cuvette 5 = 5 = 5 cm3 . was used to isolate the drop from its environment. It has been found that there are virtually no difference between the measured contact angles with or without a cuvette. The dynamic advancing and receding contact angle measurements can be performed, respectively, by pushing or pulling the syringe plunger of a motorized syringe mechanism, leading to an increase or decrease in drop volume. A schematic of this mechanism is shown in Fig. 4. Normally, at least five and up to 10 dynamic contact angle measurements were performed on a new solid surface each time, at velocities of the three-phase contact line in the range from 0.1 to 1.5 mmrmin. It will become apparent later that low-rate dynamic contact angles in this velocity range are essentially identical to the static contact angles for relatively smooth surfaces. In these dynamic procedures, liquid is supplied into the sessile drop from below the solid surface using a motorized-syringe device w80x. In order to facilitate such an experimental procedure, a hole of approximately 1 mm diameter in the solid surface is required. The strategy of pumping liquid from below the surface was pioneered by Oliver et al. w81,82x because of its potential for avoiding drop vibrations and for measuring true advancing contact angles without disturbing the drop profile. In order to avoid leakage between a stainless steel needle and the hole on the surface ., Teflon tape was wrapped around the end of the needle before insertion into the hole. In the literature, it is customary to first deposit a drop of liquid on a given solid surface using a syringe or a Teflon needle; the drop is then made to advance by supplying more liquid from above using a syringe or a needle in contact with the drop. Such experimental procedures cannot be used for

178

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Fig. 4. Schematic of a motorized-syringe mechanism for dynamic contact angle measurements.

ADSA-P since ADSA determines the contact angles and surface tensions based on the complete and undisturbed drop profile. In actual experiments, an initial liquid drop of approximately 0.3 cm radius was carefully deposited from above using a gas-tight Hamilton syringe with a stainless steel needle, covering the hole in the surface. This is to ensure that the drop will increase axisymmetrically in the centre of the image field when liquid is supplied from the bottom of the surface and will not hinge on the lip of the hole. The motor in the motorized-syringe mechanism was then set to a specific speed, by adjusting the voltage from a voltage controller. Such a syringe mechanism pushes the syringe plunger, leading to an increase in drop volume and hence the three-phase contact radius. A sequence of pictures of the growing drop was then recorded by the computer typically at a rate of one picture every 2 s, until the three-phase contact radius was approximately 0.5 cm or larger. For each low-rate dynamic contact angle experiment, at least 50 and up to 500 images were normally taken. Since ADSA-P determines the contact angle and the three-phase contact radius simultaneously for each image, the advancing dynamic contact angles as a function of the three-phase contact radius i.e. location on the surface . can be obtained. In addition, the change in the contact angle, drop volume, drop surface area, and the three-phase contact radius can also be studied as a function of time. The actual rate of advancing can be determined by linear regression, by plotting the three-phase contact radius over time. For each liquid, different rates of advancing were studied, by adjusting the speed of the pumping mechanism w80x.

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

179

It should be noted that measuring contact angles as a function of the three-phase contact radius has an additional advantage: the quality of the surface is observed indirectly in the measured contact angles. If a solid surface is not very smooth, irregular and inconsistent contact angle values will be seen as a function of the three-phase contact radius. When the measured contact angles are essentially constant as a function of surface location, the mean contact angle for a specific rate of advancing can be obtained by averaging the contact angles, after the three-phase contact radius reaches 0.3 0.5 cm see later.. The purpose of choosing these relatively large drops is to avoid any possible line tension effects on the measured contact angles w59,83,84x. 2.3.2. Inert (non-polar) surfaces: FC-722-coated mica surface A fluorocarbon, FC-722, dip-coated onto mica surfaces was chosen as the substrate for the dynamic contact angle experiments. FC-722 is a fluorochemical coating available from 3M and is chemically very similar to the FC-721 used in earlier studies. They were prepared by the same dip-coating procedures described elsewhere w64,65x, using freshly cleaved mica surfaces. The mica surfaces were obtained originally as sheets. Before dip-coating, the mica surfaces were prepared in the following procedures: 1. mica sheets were cut into small mica plates of approximately 2 = 3 cm2 ; 2. a hole of approximately 1 mm in diameter was made, by drilling, in the centre of each mica surface; 3. each mica surface was then cleaved by a sharp knife, cleaned with ethanol, and acetone, and dried in the air before dip-coating. Seventeen liquids were chosen for the contact angle measurements w80x. The surface tensions of these liquids were determined independently by applying ADSA-P to pendant drops at room temperature, 23.0 " 0.5 C. Selection of these liquids was based on the criteria described in the preceding section. Fig. 5 shows a typical example of a low-rate dynamic contact angle experiment: water on a FC-722 surface. As can be seen in this figure, increasing the drop volume V linearly from 0.18 cm3 to 0.22 cm3, by the motorized-syringe mechanism, increases the contact angle from approximately 108 to 119 at essentially constant three-phase contact radius R. This is due to the fact that even carefully putting an initial drop from above on a solid surface can result in a contact angle somewhere between advancing and receding. Therefore, it takes addition of a certain amount of liquid for the initial drop front to start advancing. Further increase in the drop volume causes the three-phase contact line to advance, with essentially constant as R increases. Increasing the drop volume in this manner ensures the measured to be an advancing contact angle. The rate of advancing for this experiment can be determined by linear regression from the linear region of the plot of the three-phase contact radius over time: It was found that the drop periphery was being advanced at a rate of 0.14 mmrmin, in the specific example given in Fig. 5. The regression coefficient for the rate of advancing was found to be 0.999; this indicates that the rate of change of the three-phase contact line was very constant, even though it was controlled simply by manipulating the drop volume.

180

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Fig. 5. Low-rate dynamic contact angles of water on a FC-722-coated mica w80x.

As can be seen in Fig. 5, the measured contact angles are essentially constant as R increases. This indicates good surface quality of the solid used in this experiment. It turns out that averaging the measured contact angles after R reaches 0.50 cm is convenient, since the drop is guaranteed to be in the advancing mode and that line tension effects are negligible w59,83,84x. Averaging the measured contact angles, after R reaches 0.50 cm, yields a mean contact angle of 118.4 for water. While a three-phase contact radius of 0.50 cm may seem to be an arbitrary value, it turns out that there is virtually no difference between averaging for R larger than 0.48 cm and 0.54 cm; the contact angles are essentially constant after R s 0.50 cm. Similar results were also obtained for other liquids w80x. Dynamicrstatic contact angle experiments have also been performed: A liquid drop is first selected to advance at a specific rate of advancing and then stopped, while a sequence of images is recorded. A typical experiment is shown in Fig. 6 for cis-decalin. As drop volume increases from 0.05 cm3 to 0.07 cm3, the three-phase contact line advances from approximately 0.36 cm to 0.41 cm at a rate of 0.41 mmrmin. A sequence of drop images was acquired after the motor was stopped at R s 0.41 cm. As can be seen in Fig. 6, the contact angle is independent of the slow rate of advancing, suggesting that the low-rate dynamic contact angle d y n is identical to the static contact angle st at . This result re-confirms validity of the experimental protocol used by Li et al. w64,65x to measure static contact angles and

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

181

Fig. 6. Dynamic and static contact angles of cis-decalin on a FC-722-coated mica w80x. This result suggests that the low-rate dynamic contact angle is identical to the static angle.

is also in good agreement with recent work to determine low-rate dynamic contact angles by the automated capillary rise technique w85,86x. With each liquid, ten different measurements i.e. ten different rates of advancing on ten new surfaces . were performed w80x. It was found that these dynamic contact angles are essentially independent of the velocity of the three-phase contact line, as is, in principle, obvious from Fig. 6. Since the low-rate dynamic contact angles are essentially independent of the velocity of the three-phase contact line, a mean dynamic contact angle for each pure liquid was determined by averaging the contact angles at the ten different rates of advancing. Fig. 7 shows all experimental contact angle results at very slow motion of the three-phase contact line in a plot of l cos vs. l . This result reconfirms the finding of Li et al. w64,65x shown in Fig. 2 that, at constant s , l cos changes smoothly with l , independent of the liquid properties. Clearly, if there were any additional and independent effects of intermolecular forces on the contact angles, the data points for the polar liquids 1-pentanol, 1-decanol, DMSO, 2,2 thiodiethanol, diethylene glycol, ethylene glycol, glycerol, and water. in Fig. 7 would not fall completely on a smooth curve along with the non-polar liquids. It can be seen that the data point for 1-pentanol is slightly below the curve; 1-decanol is again slightly below; DMSO is above and ethylene glycol is on the curve: there is

182

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Fig. 7. l cos vs. l for the dynamic contact angles of the 17 liquids on a FC-722-coated mica. This result reconfirms the finding of Li et al. shown in Fig. 2 that at constant s , l cos changes smoothly with l , independent of other liquid properties.

no evidence of any systematic deviation of the polar liquids away from such a curve. This question will also be addressed in Section 2.4.1. It should be noted that in these low-rate dynamic contact angle experiments, images of an advancing drop and hence information, such as surface tension and contact angle. are recorded continuously as drop volume is steadily increased from below the surface. The procedures used here are different from those by Li et al. w64,65x in that the contact angles measured by Li et al. were static angles, i.e. contact angles at zero velocity of the three-phase contact line. 2.3.3. Non-inert (polar) surfaces: poly(propene-alt-N-(n-hexyl)maleimide) and poly(propene-alt-N-(n-propyl)maleimide) Experience has shown that non-polar surfaces, such as Teflon and fluorocarbons, often are quite inert with respect to many liquids; however, polar surfaces often are less inert and hence may show different contact angle patterns, due to such causes as chemical reaction andror swelling and dissolution of the solid by the liquid. Since low-rate dynamic contact angle experiments when interpreted by ADSA-P have many advantages over the conventional way of manually putting tangents to the sessile drops, ADSA-P is employed here to measure low-rate dynamic contact angles on the copolymers, polypropene-alt-N- n-propyl.maleimide. and polypropene-alt-N- n-hexyl.maleimide., in order to elucidate the discrepancies between the results in Figs. 2 and 7 on the one hand, and those in Fig. 3 by a goniometer on the other. It will become apparent that a goniometer technique is liable to produce a mixture of meaningful and meaningless angles, with no criteria to distinguish between the two. If one disregards the meaningless angles to be identified by dynamic ADSA measurements, the results are in harmony with those patterns shown in Figs. 2 and 7, as will be shown below.

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

183

Eight liquids were selected for the contact angle measurements on polypropenealt-N- n-propyl.maleimide. and 13 liquids on polypropene-alt-N- n-hexyl.maleimide.. The chemical structure of these copolymers has been described elsewhere w76x. Silicon wafer 100: was again selected as the substrate. They were obtained as circular discs of approximately 10 cm diameter and were cut into rectangular size of approximately 2 = 3 cm2 . The general procedures to prepare the solid surfaces are similar to those described in Section 2.2. Before soaking the rectangular wafer surfaces into chromic acid, a hole of approximately 1 mm diameter was made by drilling, using a diamond drill bit from Lunzer New York; SMS-0.027.. For each copolymer, a 2% solution was prepared using tetrahydrofuran as the solvent. A few drops of the 2% copolymerrtetrahydrofuran solution were deposited on the dried silicon wafers inside petri glass dishes overnight; the solution spread and a thin layer of the copolymer formed on the wafer surface after tetrahydrofuran evaporated. This procedure produced good quality coated surfaces. Such polymer-coated surfaces have been studied by Atomic Force Microscopy w87x. It was found that surface roughness of the surfaces prepared by this procedure is in the order of nanometer. Fig. 8 shows a typical example of water on the polypropene-alt-N- npropyl.maleimide. copolymer. As described above, in order to avoid that the drop

Fig. 8. Low-rate dynamic contact angles of water on a polypropene-alt-N- n-propyl.maleimide. copolymer w76x.

184

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

hinges at the edge of the hole, a small drop is deposited from above, covering the hole completely. This leads to an initial contact angle different from the advancing angle: increasing the drop volume, V, linearly from 0.10 cm3 to approximately 0.12 cm3 increases the apparent contact angle, , from approximately 72 to 77 at essentially constant three-phase radius, R. This is due to the fact that even carefully putting an initial water drop from above on a solid surface can result in a contact angle somewhere between advancing and receding. This effect can be pronounced for liquids, such as water, which evaporate fast. Thus, it takes time for the initial drop front to start advancing. Further increase in the drop volume causes the three-phase contact line to advance, with essentially constant as R increases. Increasing the drop volume in this manner ensures the measured to be an advancing contact angle. In this specific example, the measured contact angles are essentially constant as R increases. This indicates good quality of the surfaces used. A mean contact angle of 77.33 " 0.06 was obtained for this experiment. A total of nine experiments at different rates of advancing were performed, each on a newly prepared surface. Since the contact angle results of water at different rates of advancing were found to be essentially constant w76x, they were averaged and resulted in a final value of 77.51 " 0.27 . Another contact angle result is shown in Fig. 9 for glycerol on the same

Fig. 9. Low-rate dynamic contact angles of glycerol on a polypropene-alt-N- n-propyl.maleimide. copolymer w76x.

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

185

copolymer. Because the contact angles are essentially constant for all experiments, they can be averaged, resulting in a mean contact angle of 70.82 " 0.05 at a rate of advancing of 0.16 mmrmin. Similar results were also obtained for 2,2 thiodiethanol and 1-bromonaphthalene w76x. For each liquid, at least 5 and up to 10 different experiments for different rates of advancing. were performed. Unfortunately, not all liquids yield essentially constant advancing contact angles. Fig. 10 shows the results of formamide on the same copolymer. It can be seen that as drop volume increases initially, contact angle increases from 60 to 63 at essentially constant three-phase radius. As the drop volume continues to increase, suddenly decreases to 60 and the three-phase contact line starts to move. As R increases further, the contact angle decreases slowly from 60 to 54 . The surface tension-time plot indicates that the surface tension of formamide decreases with time. This suggests that dissolution of the copolymer occurs, causing l and to change from that of the pure liquid. Similar behavior can also be observed in other experiments w76x. It is an important question to ask which contact angles one should use for the interpretation in terms of surface energetics. Since chemical or physical reactions, such as polymer dissolution change the liquid vapor, solid vapor and solid liquid interface interfacial tensions. in an unknown manner, the contact angle data in Fig. 10 should be disregarded for the interpretation in terms of

Fig. 10. Low-rate dynamic contact angles of formamide on a polypropene-alt-N- n-propyl.maleimide. copolymer w76x.

186

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

surface energetics. The criteria for rejecting angles for calculations of solid surface tensions will be discussed in detail in Section 2.5. Obviously, it is virtually impossible for goniometer measurements to detect the complexities shown in Fig. 10, e.g. the decrease in l . Thus, the contact angle obtained from a goniometer for this and similar solid liquid systems cannot be meaningful. The results of ethylene glycol are shown in Fig. 11. It can be seen that the contact angle increases slowly from 54 to 57 as the three-phase contact line advances from 0.46 cm to 0.54 cm. While the cause of this increase in the contact angle is unclear, it is suspected that the operative solid liquid interfacial tension is changed slowly due to physico-chemical reaction: according to Youngs equation, if the values of l and s , are constant, a change in the contact angle must be a consequence of a change in sl . It should be noted that the observed trends in the contact angle may well start immediately after drop formation, not only after the measurement procedure was set in motion. Also, there is no reason to suspect that the change of contact angle with time or radius. would cease once the measurement was terminated. More likely, such trends would continue. Because there is no unique apparent contact angle and it is unclear whether or not the solid liquid interfacial tension will remain constant and whether Youngs equation is applicable, these angles should be excluded from the interpretation in terms of surface

Fig. 11. Low-rate dynamic contact angles of ethylene glycol on a polypropene-alt-N- n-propyl.maleimide. copolymer w76x.

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

187

energetics. However, one might consider averaging the contact angles larger than R s 0.48 cm, since l seems to be constant and since the contact angle error after averaging would not be very large, i.e. s 55.74 "0.63 . It should be noted that such averaging is not allowed: apart from the experimental reasons given above, the question of whether averaging over some data is allowed or not has firm quantitative answers based on the laws of statistics. The main fact is that the statistical analysis rejects averaging of the above contact angle data over time by a very large margin w76x. While the contact angle data of formamide, diiodomethane, ethylene glycol, and diethylene glycol for polypropene-alt-N- n-propyl.maleimide. should be disregarded, the contact angles of water, glycerol, 2,2 -thiodiethanol, and 1bromonaphthalene can be used for the interpretation in terms of surface energetics. A similar study was conducted for a second copolymer, polypropene-alt-N- nhexyl.maleimide.. It was found that only the advancing contact angles of water, glycerol, diethylene glycol, and cis-decalin are essentially constant w76x. The remaining liquids, formamide, 2,2 -thiodiethanol, diiodomethane, ethylene glycol, 1-bromonaphthalene, dimethyl sulfoxide DMSO., dibenzylamine, ethyl cinnamate, and 2,5-dichlorotoluene all show very complex contact angle patterns which have to be excluded from the interpretation in terms of surface energetics and testing of approaches for interfacial tensions. A different contact angle experiment is shown in Fig. 12 for diiodomethane on the polypropene-alt-N- n-hexyl.maleimide.: initially the apparent drop volume, as perceived by ADSA-P, increases linearly, and the contact angle increases from 88 to 96 at essentially constant three-phase radius. Suddenly, the drop front jumps to a new location as more liquid is supplied into the sessile drop. The resulting contact angle decreases sharply from 96 to 88 . As more liquid is supplied into the sessile drop, the contact angle increases again. Such sliprstick behavior could be due to non-inertness of the surface. Phenomenologically, an energy barrier for the drop front exists, resulting in sticking, which causes to increase at constant R. However, as more liquid is supplied into the sessile drop, the drop front possesses enough energy to overcome the energy barrier, resulting in slipping, which causes to decrease suddenly. It should be noted that as the drop front jumps from one location to the next, it is unlikely that the drop will remain axisymmetric. Such a non-axisymmetric: drop will obviously not meet the basic assumptions underlying ADSA-P, causing possible errors, e.g. in the apparent surface tension and drop volume. This can be seen from the discontinuity of the apparent drop volume and apparent surface tension with time as the drop front sticks and slips. Similar behavior can also be observed in other experiments w76,88x. Obviously, the observed contact angles cannot all be Young contact angles, since l , s and sl . are constants, so that because of Youngs equation, ought to be a constant. In addition, it is difficult to decide unambiguously at this moment whether or not Youngs equation is applicable at all because of lack of understanding of the sliprstick mechanism. Therefore, these contact angles should not be used for the interpretation in terms of surface energetics.

188

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Fig. 12. Low-rate dynamic contact angles of diiodomethane on a polypropene-alt-N- n-hexyl.maleimide. copolymer w76x.

While pronounced cases of sliprstick behavior can indeed be observed by the goniometer, it is virtually impossible to record the entire sliprstick behavior manually. In this case, the goniometer contact angle can be very subjective, depending on the skill of the experimentalist. It is expected that a contact angle thus recorded by the goniometer should agree with the maximum angles obtained by ADSA-P. Indeed, a contact angle of 98 was observed, in reasonable agreement with the maxima in the entire sliprstick pattern of the ADSA-P results f 96 in Fig. 12.. In cases where the liquid vapor surface tension of the sessile drop decreases due to, e.g. dissolution of the surface, only ADSA can detect changes in the liquid vapor surface tension. The distinctions and differentiations made by ADSA-P are not possible in a goniometer study. Thus, circumspection is necessary in the decision whether or not experimental contact angles can be used in conjunction with Youngs equation; contact angles from a conventional goniometer sessile drop technique may produce contact angles which do not fulfil the basic assumptions made in all surface energetic approaches w10 17x, e.g. constancy of l and applicability of Youngs equation already mentioned in the Introduction. These various assumptions will be discussed in detail in Section 2.5. A comparison of the angles measured here and those by the goniometer technique had been made. It was found that the goniometer angles correspond very

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

189

Fig. 13. l cos vs. l for the polypropene-alt-N- n-propyl.maleimide. and polypropene-alt-N- nhexyl.maleimide. copolymers w76x. Fig. 3 changes drastically upon elimination of the angles shown to be meaningless using the dynamic contact angle procedures by ADSA-P.

well with those of ADSA-P only in situations where the contact angles are constant; in cases where complexities of contact angles arise, the goniometer contact angles correspond only to the maxima of the angles from ADSA-P. The picture emerging in Fig. 3 changes drastically upon elimination of the angles shown to be meaningless in the ADSA-P study, see Fig. 13. The curves in Fig. 13 are in harmony with the results obtained for more inert polar and non-polar surfaces in Figs. 2 and 7. Again, it can be concluded that l cos changes smoothly and systematically with l at constant s . Because of Youngs equation, a relation of the form of Eq. 5. can be deduced:
sl

s F

5 .

Thus, the experimental procedures and techniques used are crucial in the collection of contact angle data for surface energetics; conventional goniometer techniques may produce contact angles which violate the basic assumptions made in all surface energetic approaches. In other words, the most serious shortcoming of goniometer studies is not subjectivity and lack of accuracy, but inability to distinguish between meaningful and meaningless contact angle measurements. A recent study w89x has shown that using the above dynamic procedures with a simple polynomial fit to sessile drop images may also allow one to identify meaningless and meaningful contact angles. 2.3.4. Other non-polar and polar surfaces Similarly, low-rate dynamic contact angles of a large number of liquids were studied extensively on various solid surfaces. They are FC-722-coated silicon wafer w89x, FC-725-coated silicon wafer w90x, poly n-butyl methacrylate . PnBMA w91x,

190

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

polystyrene PS w92x, polystyrene-alt-hexylr10-carboxydecyl90r10.maleimide.. w93x, polymethyl methacrylatern-butyl methacrylate. PMMArnBMA. w94x, polymethyl methacrylate . PMMA w95x, polypropene-alt-N-methylmaleimide. w89x. Details with respect to solid surface preparation and experimental results have been given elsewhere w89 95x. Again, it was found that not all contact angles can be used for energetic purposes. For example, sliprstick of the three-phase contact line, l decreases, and decreases or increases as the drop front advances, and dissolution of the polymers by the liquids were observed. The meaningful angles for these polymers

Fig. 14. l cos vs. l for a. FC-725-coated. silicon wafer w90x, b. poly n-butyl methacrylate . PnBMA w91x, c. polystyrene PS w92x, d. polystyrene-alt-hexylr10-carboxydecyl90r10.maleimide.. w93x, e. polymethyl methacrylatern-butyl methacrylate . PMMArnBMA. w94x, f. polymethyl methacrylate . PMMA w95x, and g. polypropene-alt-N- n-alkyl.maleimide. w89x.

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

191

Fig. 14. Continued

copolymers. are plotted in Fig. 14, in plots of l cos vs. l . For a given solid surface, s is expected to be constant and these results suggest that l cos changes smoothly and systematically with l . Changing the solid surface and hence s . shifts the curves in a regular manner: intermolecular forces do not have any additional and independent effects on the contact angles. The universality of these contact angle patterns will be discussed in the next section. 2.4. Uni ersality of contact angle patterns In addition to ADSA-P, the capillary rise at a vertical plate technique w85,86,96x is also suitable for the determination of low-rate dynamic contact angles. Since the capillary rise at a vertical plate technique has been automated w96x to perform various dynamic advancing and receding contact angle measurements at immersion

192

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Fig. 14. Continued

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

193

speeds ranging from 0.008 to 0.9 mmrmin, measurements can be and have been performed w85,86x under dynamic conditions closely resembling those in the ADSAP experiments. The capillary rise experiments are performed by immersing a vertical plate. solid surface into the liquid, at a constant slow. speed; during the immersion, motion of the three-phase contact line along the vertical solid surface is always monitored. The task of measuring a contact angle is reduced to the measurement of a length capillary rise., which can be performed optically with a high degree of accuracy. In several instances, the capillary rise technique has been employed on relatively inert and well-prepared surfaces: FC-721-coated mica w86x, heat-pressed Teflon FEP w86x, hexatriacontane w51,97x, and cholesteryl acetate w51,98x. The FC-721 and FEP surfaces were prepared, respectively, by a dip-coating technique and a heat-pressing technique w64x. Hexatriacontane and cholesteryl acetate were produced by vapor deposition in a vacuum w51,97,98x; the surface quality was found to be so good that there was no contact angle hysteresis when water was used. It has been found that the contact angles determined from ADSA-P and the capillary rise at a vertical plate technique are virtually identical for the same solid liquid systems w85,86x.

Fig. 15.

l cos

vs.

for various solid surfaces.

194

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

In order to illustrate the universality of the contact angle patterns, Fig. 15 shows recent dynamic contact angles for various solids measured by the capillary rise technique and those by ADSA-P: FC-721-coated mica w86x, FC-722-coated mica w80x and silicon wafer w89x, FC-725-coated silicon wafer w90x, heat-pressed Teflon FEP w86x, hexatriacontane w51,97x, cholesteryl acetate w51,98x, polypropene-alt-N n-hexyl.maleimide. w76,88x, poly n-butyl methacrylate . PnBMA w91x, polystyrene PS w92 x, poly styrene-alt- hexylr10-carboxydecyl 90r10 .maleimide .. w93 x, polymethylmethacrylatern-butyl methacrylate . PMMArnBMA. w94x, polypropene-alt-N- n-propyl.maleimide. w76,88x, polymethyl methacrylate . PMMA w95x and polypropene-alt-N-methylmaleimide. w89x. Details of the solid surface preparation can be found elsewhere w51,76,80,86,88 95,97,98x. The ADSA-P low-rate. dynamic contact angles were established also by the procedures developed in the preceding section. Again, the results in Fig. 15 suggest that the values of l cos change systematically with l in a very regular fashion, from the hydrophobic hydrocarbon surfaces to the hydrophilic polypropene-alt-N-methylmaleimide. surface, and that the patterns are independent of the experimental technique on the one hand and liquid structure on the other. Overall, the regularity of the contact angle patterns is remarkable. For a given solid surface, s is assumed reasonably to be constant; the results suggest that l cos depends only on l . Changing the solid surface and hence s . shifts the curve in a very regular manner. Overall, these experimental results suggest that l cos depends only on l and s wsee Eq. 3.x where f is as yet an unknown function. The specific intermolecular forces of the liquids and solids, which give rise to the surface tensions, do not have additional and independent effects on the contact angles. Thus, one can change the contact angle, and because of Youngs equation, the solid liquid interfacial tension, by the simple mechanism of changing either the liquid or the solid. Combining Eq. 3. with Youngs equation, we can express sl as another unknown function F of only l and s , allowing a search for an equation in the form of Eq. 5.. This is called an equation-of-state relation, as a relation involving intensive properties w99x. It should be noted that equations of this form have been in the literature for a long time, e.g. Antonows w100x and Berthelots w101x rules see later.. 2.4.1. Reasons of de iation from smoothness As Fig. 15 stands, on one and the same solid surface, there are minor deviations of some contact angle data from the curves and one might wish to argue that intermolecular forces could still have some independent effects on the contact angles. This question will be addressed here, by focusing on the contact angle data of three chemically similar methacrylate polymers. Fig. 16 shows a plot of l cos vs. l curves for the PnBMA, PMMArnBMA., and PMMA polymers. As can be seen in this figure, there are slight deviations of the data points from the curves; however, it is not apparent that these deviations are systematic and are caused by any independent effects of intermolecular forces on the contact angles: For example, in the case of 3-pyridylcarbinol with l of approximately 48.0 mJrm2 , the data point on the PMMA curve appears to be

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

195

Fig. 16.

l cos

vs.

for three methacrylate polymer-coated silicon wafer surfaces.

slightly higher, while the point on the PMMArnBMA. is slightly lower and the one for PnBMA slightly higher again. There is no evidence for any systematic variation, for this and other liquids. Clearly, if there were a deviation due to specific intermolecular forces, one would, at the most, expect a monotonous change of the deviation when going from PMMA to PMMArnBMA. to PnBMA. Obviously, this does not mean that intermolecular forces are irrelevant; they determine the primary surface tensions of the liquids and solids. To keep matters in perspective, it has to be realized that the curves in Fig. 16 would have to be considered completely smooth if a conventional goniometer technique with "2 contact angle accuracy had been used. Furthermore, the fact that an equilibrium spreading pressure e
e

9.

of as low as 1 mJrm2 would easily contribute to such a variation should not be overlooked; s is the surface tension of a solid in a vacuum. On one and the same solid surface, s is expected to be constant when vapor adsorption is negligible, however, if vapor adsorption of liquids does play a role, then s , can be different from liquid to liquid even on one and the same type of solid surface. The answer to the question of whether or not vapor adsorption has an effect on the contact angle patterns in Fig. 16 and similar figures. would require general criteria to distinguish this from all other effects, such as swelling of a solid surface. As Fig. 16 stands, it does not appear that vapor adsorption plays a significant role in the contact angles presented here; otherwise, one would expect such an effect to manifest itself as random variations of the contact angles, rather than the remarkable regularity. Nevertheless, minor adsorption which causes the equilibrium spreading pressure 2 e , to vary by as much as 1 mJrm has been considered possible for low-energy solid surfaces w102x.

196

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Although the liquids and solids polymers. selected are of high purity, e.g. ) 99%, minute impurities presented in either the liquids or the solids are unavoidable and finding liquids and solids polymers. which contain absolutely no impurities is unrealistic. There is no guarantee that such matters could not have caused the apparent minor deviations of the points in Fig. 16 and similar figures.. It is worth mentioning that even very minor swelling of the polymer w54x or creeping of the liquid could easily introduce a slight deviation of some points away from such curves. Therefore, only after considering all such possibilities would the need or justification arise for explanations in terms of direct effects of intermolecular forces w45x. There is no reason to suppose that intermolecular forces have any additional and independent effects on the contact angles beyond the fact that intermolecular forces determine the primary interfacial tensions, l , s and sl .; the interfacial tensions then determine the contact angle, as given by Youngs equation. This fact is most easily understood in terms of the experimental contact angle patterns shown in Figs. 2, 7 and 13 16. 2.5. Criteria for calculations of surface energetics In Section 2.3.3, it has been shown that there exist a large number of contact angle complexities which prevent use of the measurements for energetic calculations. To arrive at the curves shown in Fig. 15, several assumptions already mentioned before were used to eliminate the meaningless contact angles. However, there remains one last assumption which has not been used and discussed: the constancy of s , going from liquid to liquid. It will be apparent in this section that such an assumption is indeed needed for energetic calculations; the importance of the various underlying assumptions in the determination of solid surface tensions will be discussed in detail. They will be addressed further in Section 3.1, in conjunction with the surface tension component approaches w11,15 17x. 2.5.1. Accepted assumptions for calculations of surface energetics 1. It has to be realized that Youngs equation Eq. 1. is the only thermodynamic relation to inter-relate the three surface tensions, l , s , sl , with the experimental contact angle . Therefore, the expectation of the applicability of Youngs equation has to be fulfilled. As in the case of sliprstick of the three-phase contact line shown in Fig. 12, Youngs equation cannot be applicable: Since l , s , and sl are material properties and are expected to be constant, Youngs equation implies a unique contact angle. Thus, contact angles from sliprstick of the three-phase contact line have to be discarded. 2. Obviously, contact angle interpretation of surfactant solutions or mixtures of liquids is expected to be more complicated than that of the pure liquid. It has been found that if one measures contact angles of mixture solutions on one and the same solid surface, scatter, or patterns different from those in plots of the type in Fig. 15 would arise w103x. Thermodynamically, such systems have three degrees of freedom w72 75x. The additional degree of freedom comes from the

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

197

effect of an additional liquid component. While Youngs equation may still be applicable in this case, no contact angle approach as yet allows the determination of solid surface tensions from such angles. Therefore, pure liquids are always used in contact angle measurements. However, even if this is the case, one has to ensure that l remains constant during the experiment. In an example shown earlier in Fig. 10, polymer dissolution occurs, causing the liquid surface tension to differ from that of the pure liquid. This reflects the fact that the presumed pure liquid has been changed to a mixture of liquidrpolymer solution. If a conventional goniometer technique had been used instead, the change in the operative l might not be detected. Such goniometer measurements would inevitably reflect a solid liquid system with a changed l , rather than the l of the presumed pure liquid. Thus, contact angle interpretation from such angles would be in error. 3. If a chemicalrphysical reaction takes place, any of l , s , and sl could change during the experiment, and because of Youngs Eq. 1., the contact angle would also change. Therefore, changes in suggest that at least one of l , s , and sl is changing. In an example shown in Fig. 11, the contact angle increases as the drop front advances with essentially constant l . While there is no reason to question the applicability of Youngs equation, sl is suspected to change; from Youngs equation, if l and s are constant, changes in must be a consequence of a change in sl . While such cases might be very interesting in a different context, they have to be discarded because our goal is to deduce solid surface tension from the simplest possible situations. 4. If l f s or l - s , complete wetting occurs, or a change in s induced by the solidrliquid contact autophobicity.. In order to illustrate this, a plot of vs. l is shown in Fig. 17 for a PET surface. In this figure, liquids l cos

Fig. 17.

l cos

vs.

for a PET surface.

198

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

having surface tensions less than that of the anticipated PET surface tension are included, i.e. l - s . As can be seen, the values of l cos increase as l decreases, reaching a global maximum. Further decrease in l causes the data point to fall on a straight line. Thus, the liquid surface tensions of the testing liquids should be higher than that of the anticipated solid surface tension, by the appropriate choice of the liquids. Another possible effect of l - s is liquid adsorption, which could cause s to be different from liquid to liquid. Therefore, the testing liquids used in this study were selected to fulfil the condition l ) s . 5. One last assumption which has not been used to identify meaningful contact angles in Section 2.3 is the assumption of the constancy of s going from liquid to liquid. This assumption is needed for reasons of procedure in deducing solid surface tensions. Between the three variables, l , s , and sl , only l is measurable; an additional measurable quantity is, however, the contact angle which can be inter-related to these quantities only by Youngs equation. Therefore, it is impossible to study the direct effect of changing l on a second non-measurable . quantity sl , through , unless the third non-measurable . quantity, s , is kept constant. Otherwise, interpretation of the data showing implicitly the effect of changes of l on sl would not be possible by any of the contact angle approaches of current interest. The assumption of constant s will be used to deduce solid surface tensions from contact angles in Section 3. 2.5.2. Experimental criteria The experimental results in Section 2.3.3 illustrate that there are three apparent contact angle complexities; 1. sliprstick of the three-phase contact line; 2. contact angle increasesrdecreases as the drop front advances; and 3. liquid surface tension changes as the drop front advances. These contact angles should not be used for the determination of solid surface tensions. With respect to the first point, sliprstick of the three-phase contact line indicates that Youngs equation is not applicable. Increaserdecrease in the contact angle and change in the liquid surface tension as the drop front advances violate the expectation of no physicalrchemical reaction. Therefore, when the experimental contact angles and liquid surface tensions are not constant, they should be disregarded. However, the question arises as to whether the reverse is true, i.e. whether constancy of and l in the dynamic measurements described in Section 2.3 will always guarantee that the above contact angle assumptions in the preceding section. are fulfilled. This question will be explored below. In practice, many solids are not truly inert with respect to many liquids; even swelling of fluorocarbon surfaces by alkanes has been reported w54x. For example, if swelling of a solid surface occurs quickly upon the contact with a liquid as the drop front advances, will change to reflect the changed energetics. If the time scale of such an effect is much less than that of the contact angle measurements, could reflect the changed sl . This might result in a rather constant contact angle,

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

199

although the energetics would have changed. If this mechanism would not affect the liquid, l would remain constant, i.e. be independent of time. However, such constant contact angles would reflect only the energetics of the already swollen solid surface in contact with the liquid. Furthermore, one and the same solid surface could swell differently by different test liquids in an unknown manner. This effect would manifest itself in scatter of the plots of l cos vs. l because
sl

/ F

10 .

In this case, a systematic study of the effect of surface energetics is not possible since the energetics have changed in an unknown, more complex manner. Obviously, such angles would have to be disregarded even if the measured liquid surface tensions and contact angles are constant, because the assumption of no chemicalrphysical reaction would have been violated. Unfortunately, the strategies employed here will not allow identification of such matters, and it remains a possibility that such effects contribute to the minor deviations of some data points from a smooth curve. An example for this more complex pattern is self-assembled monolayers SAMs., which have been widely used to produce monolayer surfaces of different chemical compositions and wettabilities w104 107x. For this type of surfaces, it is expected that penetration of liquids into the SAMs is inevitable; different liquids would have different effects on such surfaces. Because penetration of liquid is expected to occur almost instantly as the liquid contacts the solid, the observed advancing angle might reflect a sl value caused by a modification of the solid surface. Low-rate dynamic contact angles of water on a octadecanethiol HSCH 2 .17 CH 3 . SAM on a gold substrate are shown in Fig. 18. It can be seen that the advancing contact angles and liquid surface tension are quite constant. A plot of l cos vs. . l and of various liquids for this surface is shown in Fig. 19 solid symbols . These contact angles were determined dynamically by ADSA-P in another study w84x. Contrary to the patterns shown in Figs. 2 and 7, and Figs. 13 16, no smooth curve results in Fig. 19. The reason for the difference in pattern is believed to be the effect of liquid penetration into the monolayer SAMs. Such mechanism is believed to be absent or negligible for the relatively thick polymer films used in Section 2. It is instructive to compare this contact angle pattern with that obtained from the hexatriacontane surface w51,97,98,108x also plotted in Fig. 19.. Since both surfaces are expected to consist entirely of CH 3 groups, their solid surface tensions should be similar; one would then expect the two contact angle patterns to be essentially the same. However, the hexatriacontane data follow a smooth curve. From the point of view of surface energetics, the only difference for the two surfaces is that the SAM is a monolayer, whereas the hexatriacontane is a relatively thick crystallized layer. Obviously, it is the effect of liquid penetration and possible contact with the substrate of the SAMs which causes sl to change in a pattern different from that which would prevail in the absence of liquid penetration. In this case, there is no reason to question the validity and applicability of Youngs equation; however, different liquids are expected to penetrate

200

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Fig. 18. Low-rate dynamic contact angles of water on a octadecanethiol HSCH 2 .17 CH 3 . SAM on a gold substrate.

differently even on one and the same solid surface in an unknown manner. and hence the energetics could be very different from liquid to liquid. Since nothing is known about the changed energetics and the systems have violated the assumption of no physical reaction, use of these contact angles in any contact angle approach naively could be meaningless w109x. Although the interpretation of contact angles on SAMs in terms of surface energetics could be misleading, wettability studies of SAMs can be very interesting w107,110x.

3. Contact angle interpretation The calculation of solid surface tension s , from the contact angle of a liquid of surface tension l starts with Youngs equation, Eq. 1.. Of the four quantities in Youngs equation, only l and are readily measurable. Thus, in order to determine s , further information is necessary. Conceptually, an obvious approach is to seek one more relation among the parameters of Eq. 1., such as an equation-of-state relation, of the form of Eq. 5. which is now known to exist from experimental facts. The simultaneous solution of Eqs. 1. and 5. would solve the problem.

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

201

Fig. 19. l cos vs l for a octadecanethiol HSCH 2 .17 CH 3 . SAM on a gold substrate, and on a hexatriacontane surface.

Historically, the interpretation of contact angles in terms of solid surface energetics started with the pioneering work of Zisman and co-workers w10x; the key observation they made was that for a given solid the measured contact angles did not vary randomly as the liquid was varied; rather, cos changed smoothly with the liquid surface tension l within a band in a fashion that may suggest a straight-line relationship. Subsequent to Zismans work, two major schools of thought evolved: one is the surface tension component approach w11,15 17x; the other is the equation of state approach w12 14x which can be understood as a further development of Zismans approach. The former approach stipulates that l depends not only on l and s , but also on the specific intermolecular forces; the latter approach is based on an equation-of-state relation which has been shown to exist from thermodynamics w71 75x and the experimental results in Figs. 2, 7 and 13 16. From the experimental results shown in Figs. 2, 7 and 13 16, the curves for all solid surfaces with very different chemical properties are effectively very smooth, independent of intermolecular forces and liquid structures. This point has been addressed elsewhere w70,76,80,88 95x. Within the framework of all surface tension component approaches w11,15 17x, it stands to reason that, for a given solid surface, the values of l cos would have to depend not only on l , but also the various intermolecular forces of the liquids. This expectation, however, clashes directly with the experimental results shown in Figs. 2, 7 and 13 16. The only possible conclusion is that surface tension component approaches w11,15 17x contradict physical reality w70,76,80,88 95x. Only equation of state type approaches can be used: two well-known equation-of-state relations are those of Antonow w100x and Berthelot w101x. Nevertheless, surface tension component approaches will be briefly discussed and tested in the next section.

202

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

3.1. Surface tension component approaches 3.1.1. Fowkes approach The approach of surface tension components was pioneered by Fowkes w11x. He postulated that the total surface tension can be expressed as a sum of different surface tension components, each of which arises due to a specific type of intermolecular forces: s
d

di

11.

where , d, h , and di are, respectively, the total surface tension, dispersive surface tension component, and surface tension components due to hydrogen and dipole dipole bonding. Eq. 11. is often rearranged into s
d

12.

i.e. the total surface tension is a sum of only the dispersive d and non-dispersive n surface tension components. The former is claimed to result from molecular interaction due to London forces, the latter from all other interactions due to non-London forces. A geometric mean relationship was postulated both of the solid liquid and liquid liquid interfacial tensions:
12

y 2

d 1

d . 1r2 2

13 .

For solid liquid systems, combining Eq. 13. with Youngs Eq. 1. yields
l cos Y

s y l q 2

d s

d . 1r2 l

14 .

Typically, experimental contact angles of different liquids with known ld on a dispersive solid surface s s sd . are employed to determine the surface tension of a solid. In these procedures, the applicability of Youngs equation and the constancy of solid surface tension from liquid to liquid stated in the Section 2.5.1 are, obviously, implied. The Fowkes approach, Eq. 13. and hence Eq. 14., can be easily tested using the contact angle results in Section 2. For simplicity, two liquids of nearly the same surface tension but very different intermolecular forces are selected on the FC-722-coated mica surface; the polar non-dispersive. liquid is 1-pentanol l s 26.0 mJrm2 . and the dispersive one is trans-decalin l s ld s 27.2 mJrm2 .. From Eq. 14., the s s sd . value determined using the contact angle 73.4 . of trans-decalin is 11.2 mJrm2 . Since 1-pentanol is a polar liquid, ld is expected to be very different smaller. than the total surface tension l ; assuming for the moment that dispersion force contributes 75% to the total surface tension of 1-pentanol. This corresponds to ld s 19.5 mJrm2 for l s 26.0 mJrm2 . From the experimental contact angle 73.0 . of 1-pentanol, Eq. 14. predicts the s value to be 14.5 mJrm2 . This value is almost 30% larger than that calculated using trans-decalin; a 2 d 2 s value of 11.2 mJrm would imply l of 1-pentanol to be 25.2 mJrm . This is clearly absurd: 1-pentanol would have to be classified as a non-polar liquid, i.e.

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249
l

203

f ld. Similar calculations can be performed for other dispersivernon-dispersive liquid pairs, with similar results. This result simply illustrates the fact that the perceived surface tension components in Fowkes sense do not reflect physical reality. In principle, this is obvious from the experimental results that l cos changes smoothly with l , independent of intermolecular forces and liquid structures. Thus, the basic postulate of the Fowkes approach is false and any generalization of this approach must suffer from the same deficiency. 3.1.2. Owens Wendt Kaelble approach Owens and Wendt w15x extended Fowkes concept to cases where both dispersion and hydrogen bonding forces may operate. They regarded the surface tension as being composed of two components such that s
d

15.

where h denotes the component of surface tension due to both hydrogen bonding and dipole dipole interactions. They postulated
sl

y2

'

d s

d l

y2

'

h s

h l

16.

Combining this equation with Youngs Eq. 1. yields


l

1 q cos

.s2

'

d s

d l

q2

'

h s

h l

17 .

Nearly at the same time, Kaelble w77x also published a very similar equation in terms of dispersion and polar forces. Thus, Eq. 16. is often called the Owens Wendt Kaelble equation. Clearly, the applicability of Youngs equation discussed in Section 2.5.1 is implied in Eq. 17.. Since Eq. 17. contains two unknowns sd and sh . of the solid, it is suggested to use contact angle measurements of at least two different liquids on one and the same solid surface, by solving two simultaneous equations. Such procedures also imply constancy of the solid surface tension from liquid to liquid; if the operative solid surface tension is not constant from one liquid to the next, simultaneous solution of different equations from contact angles of different liquids. would be meaningless. Equation 17. can be easily tested using the contact angles of trans-decalin and 1-pentanol on the FC-722 surface in Fig. 7. Assuming the polarity of the dispersive. FC-722 surface to be unknown and that ld of 1-pentanol to be again 19.5 mJrm2 , two simultaneous equations can be used to determine the two unknowns sd and sh . from the experimental contact angles of 1-pentanol and trans-decalin. Eq. 17. predicts sd and sh to be, respectively, 11.2 mJrm2 and 0.6 mJrm2 . Since diiodomethane is claimed to be a dispersive liquid w16,17x l s ld s 50.0 mJrm2 . and water is a non-dispersive liquid w11x l s 72.7 mJrm2 and ld s 21.8 mJrm2 ., two additional simultaneous equations can be obtained by insertion of the experimental contact angles from Fig. 7. This procedure yields sd and sh to be 8.1 mJrm2 and 2.8 mJrm2 , respectively. Clearly, different choices of the liquid pairs

204

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

yield different surface tension components of the solid surface: if nothing were known about the dispersive property of the FC-722 surface, Eq. 17. would have predicted this dispersive surface to be non-dispersive; the relative magnitudes would depend on the choice of the liquid pairs. Similar calculations can be performed for other pairs of liquids, with similar results. 3.1.3. Lifshitz an der Waals r acid base ( an Oss) approach The Lifshitz van der Waals r acid base van Oss. approach w16,17x was claimed to be a generalization of the Fowkes approach, by considering perceived acid base interactions at the interface. van Oss et al. divided the surface tension into different perceived components, i.e. the so-called Lifshitz van der Waals LW ., acid q., and base y. components, such that the total surface tension is given by
i

LW i

q2

'

q y i i

18.

where i denotes either the solid or liquid phase. The interfacial tension was postulated both of solid liquid and liquid liquid systems as
12

y 2

LW 1

LW . 1r2 2

y 2

q y . 1r2 1 2

y 2

y q . 1r2 1 2

19 .

For solid liquid systems, combining Eq. 19. with Youngs equation yields
l

1 q cos

. s 2

LW LW . 1r2 l s

q 2

q y . 1r2 l s

q 2

y q . 1r2 l s

20.

Eq. 20. is often used to determine the solid surface tension components sLW , q, s and y. from contact angles, using three simultaneous equations by inserting s properties of calibration liquids. These procedures also imply the applicability of Youngs equation and constancy of the solid surface tension from one liquid to the next, as discussed in Section 2.5.1. The failure of this approach can be easily illustrated, by employing the postulated liquid surface tension components from van Oss et al. w16,17x and experimental contact angles of liquid triplets; three simultaneous equations of Eq. 20. from contact angles of three different polar liquids. will yield the three unknown solid properties. Table 1 illustrates the solid surface tension components of the FC-721-coated mica surface w111x. It can be seen that the calculated s , value varies from y30.0 to 107.0 mJrm2 , depending on the choice of the liquid triplets; the resulting surface tension components vary by a very large margin. However, it was recently suggested w112x that at least one non-polar liquid should be used for the above calculations, such as 1bromonaphthalene. It is the contention of the proponents of this acid base approach that 1-bromonaphthalene is a non-polar compound. No reason was given for the necessity of including a non-polar liquid. For this reason, the solid surface tension components of the same surface were recalculated using 1-bromonaphthalene and shown in Table 2. Again, the calculated s values are not consistent, varying from y36.0 to 9.5 mJrm2 . If nothing were known about the properties of the non-polar FC-721 surface, and if only a small number of measurements were available, this approach would predict all kinds of different molecular properties:

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

205

Table 1 Calculated solid surface tension components for FC-721 using three simultaneous equations of the acid base approach, Eq. 20., from contact angles of three polar liquids w111x sLW .1r2 mJrm2 . Fo-Gl-Wa Fo-Gl-DM Fo-Gl-EG Fo-Wa-DM Fo-Wa-EG Fo-DM-EG Gl-Wa-DM Gl-Wa-EG Gl-DM-EG Wa-DM-EG 2.56 2.33 2.08 7.08 1.14 1.66 3.38 9.96 0.87 2.39 sq .1r2 mJrm2 . 0.46 1.30 2.17 y4.07 1.90 2.05 y0.10 y4.72 3.23 0.97 sy .1r2 mJrm2 . 0.87 y1.60 y4.23 1.24 0.75 y2.00 0.68 y0.78 y4.76 0.53
s

mJrm2 . 7.38 1.27 y14.02 40.12 4.15 y5.44 11.31 106.6 y30.0 6.73

DM: dimethyl sulfoxide DMSO.; EG: ethylene glycol; FO: formamide; Gl: glycerol; Wa: water.

non-polar, monopolar, bipolar acidic and basic as well as negative solid and solid liquid interfacial tensions. Thus, the ability of the approach to explain, e.g. negative interfacial tensions does not imply its validity. The reason that the scatter of the results is not as bad as those in Table 1 is due to the use of Berthelots rule see later.. Clearly, this approach is just as erroneous as the earlier surface tension component approaches. 3.2. Equation of state approach In this, and the following sections, an equation of state approach will be
Table 2 Calculated solid surface tension components for FC-721 using three simultaneous equations of the acid base approach, Eq. 20., from contact angles of one non-polar and two polar liquids w111x sLW .1r2 mJrm2 . Fo-Gl-Br Fo-Wa-Br Fo-DM-Br Fo-EG-Br Gl-Wa-Br Gl-DM-Br Gl-EG-Br Wa-DM-Br Wa-EG-Br DM-EG-Br 9.07 9.07 9.07 9.07 9.07 9.07 9.07 9.07 9.07 9.07 sq .1r2 mJrm2 . y1.10 0.02 0.53 2.44 1.60 0.39 1.36 0.30 0.50 0.04 sy .1r2 mJrm2 . 5.58 0.90 y1.19 y9.16 0.76 y0.13 y3.82 0.63 0.42 2.70
s

mJrm2 . y3.21 9.11 7.81 y35.63 9.31 8.97 y 1.32 9.45 9.49 9.29

DM: dimethyl sulfoxide DMSO.; EG: ethylene glycol; FO: formamide; Gl: glycerol; Wa: water, Br: 1-bromonaphthalene.

206

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

discussed, as such a relation has been inferred from the experimental results in Section 2. In the literature, two well-known equation-of-state relations are Antonows w100x and Berthelots w101x rules. 3.2.1. Antonows rule An old equation of state is that of Antonow w100x. Antonows rule relates with l and s for solid liquid systems in the following simple manner:
sl

sl

s<

<

21 .

This equation, although never having been derived in any fashion, appears in the literature from time to time. Combining this relation with Youngs equation gives cos
Y

s y1 q 2

s l

22.

Thus, assuming, for the sake of the argument, the validity of Eq. 22., s can be determined when l and Y are known. Once s is known, sl can be found either from Youngs Eq. 1. or from Eq. 21.. 3.2.2. Berthelots (geometric mean) combining rule Another equation-of-state relation can be obtained from the Berthelot combining rule w101x. Unlike Antonows rule, Berthelots rule has a theoretical background; based on molecular interactions of like pairs; from the London theory of dispersion w113,114x, the long range dispersion energy function for two identical molecules is given by Udisp s C6 r6 23.

where C6 is a negative constant, so that the energy contribution is attractive. A more general form w114x of the long range dispersion energy function including electrostatic and induction interactions can be expressed as Udisp s C6 r
6

C8 r
8

C10 r 10

24.

where C8 and C10 are all negative constants. It has been shown w113,114x that the energy coefficient C6 for two identical molecules can be expressed as C6 s y 3 EI 4 4
2

0. .2

25.

where o is the permittivity of free space, 0. is the static polarizability, and EI , is the ionization energy. From Eq. 25., the dispersion energy coefficient of a pair of molecules of species

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

207

i is given by
ii C6

sy

3 EI i 4 4

2 . i 0

.2 o

26.

and similarly, for a pair of molecules of species j, C6j j sy 3 EI j 4 4


2 . j 0 o

.2

27 .

i It has been shown w114x that the interaction C6 j between two dissimilar molecules i and j can be related by the following expression: i C6 j s y

EI i EI j

0.

j o

0. .2

2 EI i q EI j

28.

Because the ionization energies vary only slightly from molecule to molecule, i.e. i EI i f EI j , the dispersion energy coefficient C6 j of Eq. 28. can be written in terms ii jj of C6 and C6 as
i C6 j s

'C

ii j j 6 C6

29.

This relation indeed forms a basis of the Berthelot geometric mean. combining rule w101x:
ij

'

ii

jj

30 .

where i j is the potential energy parameter well depth. of unlike-pair interactions; . ii and j j are the potential energy parameters well depth of like-pair interactions. Thermodynamically, a relation of the free energy of adhesion per unit area of a solid liquid pair is equal to the work required to separate unit area of the solid liquid interface w115x: Wsl s
l

sl

31.

From the geometric mean combining rule i.e. the Berthelot rule, Eq. 30.. the free energy of adhesion Wsl can be approximated in terms of the free energy of cohesion of the solid, Ws s , and the free energy of cohesion of the liquid, Wl l w101,114,116,117x, i.e. Wsl s Wl l Ws s

'

32 .
l

By the definitions Wl l s 2 Wsl s 2

and Ws s s 2

s ,

Eq. 32. becomes 33 .


sl

'

l s

By combining Eq. 33. with Eq. 31., the solid liquid interfacial tension

can be

208

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Table 3 Calculated s values mJrm2 . of the FC-722-coated mica surface from Antonows and Berthelots rules. The contact angles are low-rate dynamic angles measured by ADSA-P Liquid
l s

mJrm2 .

degree.

Antonows rule, Eq. 22. 16.5 16.8 17.5 17.2 17.3 19.1 19.7 20.1 21.0 20.7 20.7 20.5 20.1 21.1 20.2 20.5 18.9

Berthelots rule, Eq. 35. 11.5 10.9 11.2 10.7 10.3 11.3 10.5 10.0 10.3 9.7 9.6 8.9 8.1 7.9 6.9 6.5 4.9

Decane 1-Pentanol trans-Decalin Hexadecane 1-Decanol cis-Decalin Ethyl cinnamate Dibenzylamine Dimethyl sulfoxide DMSO. 1-Bromonaphthalene Diethylene glycol Ethylene glycol Diiodomethane 2,2 -Thiodiethanol Formamide Glycerol Water

23.88 26.01 27.19 27.62 28.99 32.32 37.17 40.80 42.68 44.31 44.68 47.55 49.98 56.26 59.08 65.02 72.70

67.36 72.95 73.38 75.94 78.84 79.56 86.54 90.70 90.95 93.81 94.22 97.87 101.18 104.56 108.49 111.73 118.69

written as
sl

y2

'

l s

'

' /
s

34.

Combining this equation with Youngs Eq. 1. yields cos s y1 q 2

s l

35.

Thus, the solid vapor surface tension can be determined when experimental Young. contact angle and liquid vapor surface tension are known. The question then arises as to what to expect when such an equation-of-state relation, in conjunction with Youngs equation, is used to calculate solid surface tensions. There is one immediate criterion that the results obtained with this equation or, for that matter, any other equation. must satisfy; when measuring contact angles with a number of liquids on a low-energy solid, the solid vapor surface tension s is expected to be constant, independent of the liquid surface tension l . In other words, different pairs of and l for one and the same solid should yield sensibly constant values of s . The s values obtained from Eq. 35. and Eq. 22. are given in Table 3 for the FC-722-coated mica from Fig. 7. It can be seen that the s values obtained from Antonows and Berthelots rule are not

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

209

constant; rather, the former tends to increase as l increases; the latter tends to decrease as l increases. As mentioned above, Antonows rule does not have a theoretical basis. But, in the theory of intermolecular interactions and the theory of mixtures, combining rules are used to evaluate the parameters of unlike-pair interactions in terms of those of the like interactions w118,119x. However, as for many other combining rules, the Berthelot rule w101x, i.e. Eq. 34., is only a useful approximation and does not provide a secure basis for the understanding of unlike-pair interactions. For the interactions between two very dissimilar types of molecules or materials, where there is an apparent difference between ii and j j , it has been demonstrated w44,120x that the geometric mean combining rule generally over-estimates the strength of the unlike-pair interaction, i.e. the geometric mean value is too large an estimate. Clearly, this is why the geometric mean combining rule does not work for situations of large differences < Wl l y Ws s < or < l y s <, as can be seen in Table 3. This fact is indeed well-known w46 49,114x. In general, sl , the minimum of the solid liquid interaction potential, is often expressed in the following manner w45 49x:
sl

s g sr l .

'

ss

ll

36 .

where s s and l l are, respectively, the minima in the solid solid and liquid liquid potentials; s and l , are the appropriate core diameters for the molecules of the solid and liquid, respectively. It should be noted that many explicit forms of g sr l . have been suggested. For example, Matyushov and Schmid w49x proposed g sr l . s

1 4

4 sr 1 q

3 l

sr l

.2

37.

and Sullivan w48x set


2

g sr l . s

1q

s l

38.

in attempts to give a better representation of solid liquid interactions from solid solid and liquid liquid interactions. In general, these functions are normalized such that g sr l . s 1 when s s l . In this case, they revert to Berthelots geometric mean combining rule, Eq. 30.. Therefore, a modification of Berthelots rule is called for, leading to an equation of state approach for solid liquid interfacial tensions. 3.2.3. Equation of state approach: modified Berthelots rule In the study of mixtures w118,119x it has become common practice to introduce a factor 1 y K i j . to the geometric mean combining rule,
ij

s 1 y K i j .

'

ii

jj

39 .

210

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

where K i j is an empirical parameter quantifying deviations from the geometric mean combining rule. Since the geometric mean combining rule overestimates the strength of the unlike-pair interactions, the modifying factor 1 y K i j . should be a decreasing function of the difference ii y j j and be equal to unity when the difference ii y j j is zero. On the basis of this thought, Li et al. w121x have considered a modified combining rule of the form
ij

'

ii

jj

ey

i iy j j .

40 .

where is an empirical constant; the square of the difference ii y j j ., rather than the difference itself, reflects the symmetry of this combining rule, and hence the anticipated symmetry of the equation of state. Correspondingly, for the cases of large differences < Wl l y Ws s < or < l y s <, the combining rule for the free energy of adhesion of a solid liquid pair can be written as Wsl s Wl l Ws s ey

'

W l lyW s s . 2

41 .
l ,

or, more explicitly, by using Wl l s 2 Wsl s 2

and Ws s s 2

s ,

'

l s

ey

l y s .

42 .

In the above, and are as yet unknown constants. Clearly, when the values of . . l and s are close to each other, Eq. 42 reverts to Eq. 33 , the geometric mean combining rule. By combining Eq. 42. with Eq. 31., an equation of state for solid liquid interfacial tensions can be written as
sl

y2

'

l s

ey

l y s .

43.

Combining Eq. 43. with Youngs Eq. 1. yields cos


Y

s y1 q 2

s l

ey

l y s .

44 .

Thus, the solid surface tensions can be determined from experimental Young. contact angles and liquid surface tensions when is known, e.g. by the Newton method w122x. Obviously, for a given set of l and data measured on one and the same type of solid surface, the constant and s values can be determined by a least-square analysis technique w64,123,124x. Starting out with arbitrary values for values s and , iterative procedures can be used to identify that pair of s and which provides the best fit of the experimental data to the set of experimental l and Y pairs belonging to one and the same solid surface. From the experimental contact angles on the FC-721-coated mica, heat-pressed Teflon FEP, and polyethylene terephthalate . PET surfaces not used here., an averaged value of 0.0001247 m2rmJ. 2 was obtained w14,64x.

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

211

Table 4 Summary of the dynamic contact angles of various solids by ADSA-P and capillary rise techniques. The 2. were calculated from the equation of state approach s values mJrm Solid surfacertechnique Liquid mJrm2 .
l s

degree.

Equation of state, Eq. 44. 11.7 11.3 11.6 11.5 11.3 12.0 11.5 11.3 11.9 11.5 11.9 11.4 11.2 12.4 12.2 12.2 12.9 12.4 12.4 12.1 11.7 12.7 12.0 12.8 12.2 12.4 11.2 11.5 12.7 11.8 12.1 11.9 12.5 11.6 13.2 12.7 12.7 11.9

FC-721-coated micar capillary rise w86x

Dodecane 2-Octanol Tetradecane 1-Octanol Hexadecane 1-Hexadecene 1-Decanol 1-Dodecanol Decane 1-Pentanol trans-Decalin Hexadecane 1-Decanol cis-Decalin Ethyl cinnamate Dibenzylamine Dimethyl sulfoxide DMSO. 1-Bromonaphthalene Diethylene glycol Ethylene glycol Diiodomethane 2,2 -Thiodiethanol Formamide Glycerol Water Hexane 2-Octanol Hexadecane Glycerol Dodecane Hexadecane 3,3-Thiodipropanol Diethylene glycol Ethylene glycol 2,2 -Thiodiethanol Formamide Glycerol Water

25.03 26.00 26.50 27.28 27.31 27.75 28.29 29.53 23.88 26.01 27.19 27.62 28.99 32.32 37.17 40.80 42.68 44.31 44.68 47.55 49.98 56.26 59.08 65.02 72.70 18.50 26.42 27.62 65.02 25.64 27.62 39.83 45.16 48.66 53.77 59.08 63.13 72.70

70.4 73.5 73.5 75.1 75.6 74.0 76.6 79.2 67.36 72.95 73.38 75.94 78.84 79.56 86.54 90.70 90.95 93.81 94.22 97.87 101.18 104.56 108.49 111.73 118.69 50.83 74.74 75.64 111.89 71.02 73.41 90.48 94.47 100.05 101.07 106.89 110.21 119.31

FC-722-coated micar ADSA-P w80x

FC-722-coated silicon waferrADSA-P w89x

FC-725-coated silicon waferrADSA-P w90x

212

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Table 4 Continued. Solid surfacertechnique Liquid


l s

mJrm2 . Teflon FEP.rcapillary rise w86x

degree.

Equation of state, Eq. 44. 17.7 17.2 17.5 17.5 17.7 17.9 18.6 17.7 18.4 20.3 20.3 20.6 20.3 21.3 21.3 21.3 21.1 28.5 28.3 26.7 28.6 27.8 28.0 29.2 29.4 28.5 29.0 28.7 29.7 30.5 29.3 29.4 30.0 30.2 31.3 30.2 31.4 31.0 31.0

Dodecane 2-Octanol Tetradecane 1-Octanol Hexadecane 1-Hexadecene 1-Dodecanol Dimethylformamide Methyl salicylate Ethylene glycol 2,2 -Thiodiethanol Glycerol Water Ethylene glycol 2,2 -Thiodiethanol Glycerol Water cis-Decalin Triacetin Diethylene glycol Glycerol Water Diethylene glycol 3-Pyridylcarbinol 2,2 -Thiodiethanol Formamide Glycerol Water Dimethyl sulfoxide DMSO. Diethylene glycol Ethylene glycol Formamide Glycerol Water Diethylene glycol Ethylene glycol Formamide Glycerol Water

25.03 26.00 26.50 27.28 27.31 27.75 29.53 35.21 38.85 47.7 54.0 63.4 72.8 47.7 54.0 63.4 72.8 32.32 35.52 44.68 65.02 72.70 45.16 47.81 53.77 59.08 65.02 72.70 42.68 44.68 48.66 59.08 63.11 72.70 45.16 48.66 58.45 63.13 72.70

47.8 52.3 52.6 54.4 53.9 54.2 55.7 68.6 72.2 79.2 86.3 95.4 104.6 77.0 84.3 94.0 103.3 28.81 39.45 61.04 82.83 92.26 58.73 60.30 68.00 76.41 82.11 90.73 50.67 52.41 61.20 74.76 78.38 88.42 51.32 59.72 70.28 76.51 87.13

Hexatriacontanercapillary rise w51,97x

Cholesteryl acetater capillary rise w51,98x

Polypropene-alt-N- nhexyl.maleimide.r ADSA-P w76,88x

Poly n-butyl methacrylate . rADSA-P w91x

PolystyrenerADSA-P w92x

Polystyrene-alt-hexylr 10-carboxydecyl90r10. maleimide..r ADSA-P w93x

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249 Table 4 Continued. Solid surfacertechnique Liquid
l s

213

mJrm2 .

degree.

Equation of state, Eq. 44. 35.7 34.2 34.6 34.0 33.3 34.6 35.9 38.6 37.1 36.5 35.7 37.0 38.3 38.4 39.0 38.3 38.6 37.9 39.3 43.7 41.7 41.8

Polymethyl methacrylater n-butyl methacrylate .r ADSA-P w94x

1-Iodonaphthalene 3-Pyridylcarbinol 2,2 -Thiodiethanol Formamide Glycerol Water 1-Iodonaphthalene 1-Bromonaphthalene 1,3-Diiodopropane 2,2 -Thiodiethanol Glycerol Water 1,3-Diiodopropane 3-Pyridylcarbinol Diiodomethane 2,2 -Thiodiethanol Formamide Glycerol Water Diiodomethane Glycerol Water

42.92 47.81 53.77 59.08 65.02 72.70 42.92 44.31 46.51 53.77 65.02 72.70 46.51 47.81 49.98 53.77 59.08 65.02 72.70 49.98 65.02 72.70

35.67 49.22 57.84 66.33 74.72 81.33 35.19 30.75 39.98 54.04 70.67 77.51 36.95 39.47 42.25 50.35 57.73 66.84 73.72 30.71 60.25 69.81

Polypropene-alt-N- npropyl.maleimide.r ADSA-P w76,88x

Polymethyl methacrylate . rADSA-P w95x

Polypropene-alt-Nmethylmaleimide. rADSA-P w89x

The s , values calculated from Eq. 44. with the above value are given in Table 4, for all solid surfaces shown in Fig. 15. It can be seen that the s values calculated from the equation of state approach w14,64x are quite constant, essentially independent of the choice of the liquids. This reconfirms the validity of the approach to determine solid surface tensions from contact angles. A FORTRAN computer program to calculate s from l and , using Eq. 44. with s 0.0001247 m2rmJ. 2 , is given in Appendix A. It should be noted that the above calculations were based on a constant value of 0.0001247 m2rmJ. 2 . Alternatively, the s values of various solids can be determined by the two-variable least-square analysis w64,123,124x, by assuming s and in Eq. 44. to be constant, as described above. This leads to a value and a s value for every given solid surface. The results, together with those obtained using the value of 0.0001247 m2rmJ. 2 , are given in Table 5. It is evident that there is good agreement between the s values determined from the two strategies, and the slight difference in the values have very little effect on the calculated s , values. Since the values do not appear to change systematically

214

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Table 5 Averaged s values reproduced from Table 4; the a two-variable least-square fit are also given Solid surfacertechnique

mJrm2 . and

m2rmJ. 2 values determined by

No. of system 8 17 4 9 9 4 4 5 6 6 5 6 6 7 3

Table 4
s a

Least-square fit
s

FC-721-coated micarcapillary rise w86x FC-722-coated micarADSA-P w80x FC-722-coated silicon waferrADSA-P w89x FC-725-coated silicon waferrADSA-P w90x Teflon FEPrcapillary rise w86x Hexatriacontanercapillary rise w51,97x Cholesteryl acetatercapillary rise w51,98x Polypropene-alt-N- n-hexyl.maleimide.rADSA-P w76,88x Poly n-butyl methacrylate .rADSA-P w91x PolystyrenerADSA-P w92x Polystyrene-alt-hexylr10-carboxydecyl90r10.maleimide..r ADSA-P w93x Polymethyl methacrylatern-butyl methacrylate .rADSA-P w94x Polypropene-alt-N- n-propyl.maleimide.rADSA-P w76,88x Polymethyl methacrylate .rADSA-P w95x Polypropene-alt-N-methylmaleimide.rADSA-P w89x
a

11.5 " 0.2 12.1 " 0.3 12.0 " 1.1 12.3 " 0.4 17.8 " 0.3 20.4 " 0.2 21.3 " 0.2 28.0 " 1.0 28.8 " 0.5 29.9 " 0.5 31.0 " 0.6 34.4 " 0.8 36.8 " 1.1 38.5 " 0.4 42.4 " 2.8

0.000124 0.000111 0.000111 0.000114 0.000142 0.000124 0.000128 0.000122 0.000124 0.000120 0.000120 0.000136 0.000133 0.000113 0.000167

11.7 11.8 11.8 11.9 18.0 20.3 21.5 27.9 28.8 29.7 30.8 34.7 36.9 38.3 43.4

Error limits are 95% confidence limits.

with the solid surface and hence s ., a weighted value can be calculated. Alternatively, a statistical analysis can be used to decide whether or not averaging of over s is allowed w76,123 126x. This procedure yields a correlation coefficient rc al of 0.477 for the 14 observations in Table 5. A tabulated correlation coefficient rt ab at 99% is 0.623. Since, rc al - rt ab , there is no correlation between and s at over 99% confidence, for the results given in Table 5. Thus, averaging of is allowed. Remarkably, this procedure together with the new experimental data yield a weighted value of 0.0001234 m2rmJ. 2 , which is in excellent agreement with that s 0.0001247 m2rmJ. 2 . obtained by Li et al. w64x from contact angles on only three solid surfaces not considered here. It should be noted that other forms of the modified Berthelots rule or any other equation in the form of Eq. 5.. can also be used, as long as they are capable of describing the experimental contact angle patterns in Figs. 2 and 7, and Figs. 13 16. A different form of the modified-Berthelots rule will be proposed and discussed in the next section, which yields essentially the same s values as Eq. 44. does. 3.2.4. Equation of state approach: alternati e formulation Following the arguments leading to Eq. 39., we propose an alternative formulation w125,126x of the equation of state approach, i.e. a modified Berthelots rule with a different modifying factor similar to the 1-K i j . used in Eq. 39.,

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

215

1y

ii

jj

.2

45 .

so that 1 y 1 ii y j j . 2 . is also a decreasing function of the difference ii y j j . and is equal to unity when i i s j j ; 1 is an unknown constant. Thus, Eq. 39. can then be written as
ij

s 1y

ii

jj

.2

/'

ii

jj

46.

The square of the difference ii y j j ., rather than the difference itself, reflects again the symmetry of this combining rule, and hence the anticipated symmetry of the equation of state. Because the free energy is directly proportional to the energy parameter w116,117x, i.e. W A , and for the cases of large differences < Wl l y Ws s < or < l y s <, the combining rule for the free energy of adhesion of a solid liquid pair can be written as Wsl s 1 y
1

Wl l y Ws s . 2 . Wl l Ws s

'

47 .
1

where 1 is an empirical constant which is in general different from more explicitly, by using Wl l s 2 l and Ws s s 2 s , Wsl s 2 1 y
1

in Eq. 46.;

.2 .

'

l s

48 .

In the above, 1 and 1 s 4 1 . are as yet unknown constants. Clearly, when the values of l and s are close to each other, Eq. 48. reverts to the geometric mean combining rule. By combining Eq. 48. with Eq. 31., an alternative equation of state for solid liquid interfacial tensions w125,126x can be written as
sl

y2

'

l s

1 y

.2 .

49.

Combining Eq. 49. with Youngs Eq. 1. yields cos


Y

s y1 q 2

s l

1 y

.2 .

50.

The question then becomes whether this combining rule fits the experimental contact angle patterns shown earlier. Obviously, for a given set of l and data measured on one and the same type of solid surface, the constant 1 and s values can be determined by a two-variable least-square analysis technique w64,123,124x. Starting out with arbitrary values for l and 1 , iterative procedures can be used to identify that pair of s and 1 values which provides the best fit of the experimental data to the set of experimental l and Y pairs belonging to one and the same solid surface. Fig. 20 shows the best-fitted curves of Eq. 50. onto the experimental contact angle data for several solid surfaces previously considered. It can be seen that Eq. 50. fits these experimental data very well. The results of such a least-square analysis are summarized in Table 6. Since the 1 values do not

216

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Table 6 mJrm2 . and m2 rmJ. 2 values determined by a two-variable least-square fit, for the two s equation-of-state formulations Solid surfacertechnique No. of Eq. 44. system 0.000124 0.000111 0.000111 0.000114 0.000142 0.000124 0.000128 0.000122 0.000124 0.000120 0.000120 0.000136 0.000133 0.000113 0.000167 Eq. 50.
s 1 s

FC-721-coated micarcapillary rise w86x 8 FC-722-coated micarADSA-P w80x 17 FC-722-coated silicon waferrADSA-P w89x 4 FC-725-coated silicon waferrADSA-P w90x 9 Teflon FEPrcapillary rise w86x 9 Hexatriacontanercapillary rise w51,97x 4 Cholesteryl acetatercapillary rise w51,98x 4 Polypropene-alt-N- n-hexyl.maleimide.rADSA-P w76,88x 5 Poly n-butyl methacrylate .rADSA-P w91x 6 PolystyrenerADSA-P w92x 6 Polystyrene-alt-hexylr10-carboxydecyl 90r10.maleimide..r 5 ADSA-P w93x Polymethyl methacrylatern-butyl methacrylate .rADSA-P w94x 6 Polypropene-alt-N- n-propyl.maleimide.rADSA-P w76,88x 6 Polymethyl methacrylate .rADSA-P w95x 7 Polypropene-alt-N-methylmaleimide.rADSA-P w89x 3

11.7 11.8 11.8 11.9 18.0 20.3 21.5 27.9 28.8 29.7 30.8 34.7 36.9 38.3 43.4

0.000097 0.000092 0.000092 0.000092 0.000118 0.000100 0.000104 0.000109 0.000108 0.000107 0.000106 0.000124 0.000123 0.000106 0.000155

11.7 11.6 11.6 11.6 17.8 19.7 20.9 27.8 28.5 29.5 30.6 34.6 36.9 38.2 43.3

appear to show any dependence on the solid surfaces w125,126x, a weighted mean .2 2 1 was calculated. This yields 1 s 0.0001057 m rmJ . A comparison between the s values for various solid surfaces obtained from the two equation of state formulations is given in Table 6. It can be seen that the s values obtained from the two formulations equations. are virtually identical. Although the two relations are very different, Eq. 45. can indeed be obtained from a first-order approximation of a Taylor-series expansion of the adjusting parameter in Eq. 40.. A FORTRAN computer program to calculate s from l and , using Eq. 50. with 2 .2 1 s 0.0001057 m rmJ , is given in Appendix B. Since Eq. 46. was formulated here as an expression similar to those of Eq. 36. and Eq. 39., insight about molecular interactions of unlike pairs can be gained; the rationale is as follows. Knowing 1 s 0.0001057 m2rmJ. 2 , the present form of the combining rule for the free energy of adhesion, Eq. 47., can be written as Wsl s 1 y 0.0000264 Wl l y Ws s .
2

. 'Wl lWs s

51 . and if W f k , a combin-

where 1 s 1r4 s 0.0000264 m2rmJ. 2 . Since W A ing rule for the energy parameter can be derived:
sl

s 1 y 0.0000264k 2

ll

ss

.2 .

'

ll

ss

52 .

where k is a proportionality constant. Although the proportionality constant k is unknown at the moment, Eq. 52.

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

217

should, in principle, be able to tell us something about the molecular interactions between unlike solid liquid pairs. In the study of molecular interactions between solids and liquids through, e.g. the Lennard Jones solid fluid potential w45,48,49,114x, a combining rule is always required to represent the interactions of unlike pairs of molecules i j from those of like pairs, ii and j j . However, the exact form of the combining rule for solid liquid energy potential is often uncertain owing to the fact that estimating solid liquid interfacial tensions directly is a difficult task and hence any specific proposed. form of the combining rule cannot be easily tested and verified. As Eq. 52. stands, information about unlike molecu-

Fig. 20. l cos vs. l for a. FC-725-coated silicon wafer w90x, b. poly n-butyl methacrylate . PnBMA w91x, c. polystyrene PS w92x, d. polystyrene-alt-hexylr10-carboxydecyl90r10.maleimide.. w93x, e. polymethyl methacrylatern-butyl methacrylate . PMMArnBMA. w94x, and f. polymethyl methacrylate. PMMA w95x. All curves are best-fits of Eq. 50. to experimental data.

218

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Fig. 20. Continued

lar interactions can be inferred when the proportionality constant k is known. It should be realized that Eq. 46. can be formulated for k to drop out, e.g. by normalization. This procedure should allow the determination of a combining rule for the specific solid liquid interactions. 3.2.5. Comparison with the original equation of state formulation The formulation of an equation of state approach is essentially an empirical curve fit to contact angle data. As such there are a variety of ways in which one might proceed. It was decided in the original formulation w12,13x to correlate the data in terms of Goods interaction parameter : s
s

q 2

sl

'

53.

l s

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

219

Fig. 20. Continued

The explicit form of such an equation-of-state relation was formulated as

sl

' / s 1 y 0.015'
l s

'

54.

l s

Combining Eq. 54. with Youngs equation yields cos s 0.015


l s

y 2.00.

'

l s

0.015'

l s

y1

55.

It should be noted that Eq. 54. and hence Eq. 55.. was derived based on

220

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Table 7 Comparison between the three equation-of-state formulations by calculating the hypothetical values of l mJrm2 . and degree.
l s

values for

mJrm2 . Kwok et al. w125x Eq. 50. 66.1 61.8 56.7 51.0 45.1 39.0 33.0 27.0 21.3 15.9 47.1 43.9 39.8 35.3 30.5 25.6 20.8 16.2 11.9 28.2 26.2 23.6 20.6 17.5 14.3 11.2 Neumann et al. w14x Eq. 55. 66.0 61.7 56.5 50.8 45.0 39.1 33.2 27.1 21.1 15.2 47.1 44.0 40.0 35.1 30.2 25.4 20.6 16.2 12.0 28.2 26.2 23.6 20.6 17.5 14.3 11.3

Li et al. w64x Eq. 44. 70.0 70.0 70.0 70.0 70.0 70.0 70.0 70.0 70.0 70.0 50.0 50.0 50.0 50.0 50.0 50.0 50.0 50.0 50.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 20.0 30.0 40.0 50.0 60.0 70.0 80.0 90.0 100.0 110.0 20.0 30.0 40.0 50.0 60.0 70.0 80.0 90.0 100.0 20.0 30.0 40.0 50.0 60.0 70.0 80.0 66.1 61.9 56.9 51.5 45.7 39.7 33.6 27.5 21.5 15.8 47.1 43.9 40.0 35.4 30.8 26.0 21.2 16.5 12.2 28.2 26.2 23.6 20.7 17.5 14.4 11.3

extensive contact angle measurements of eight polymer surfaces from Zisman et al. almost 50 years ago. This formulation was first attempted in the 1960s w12x and completed in the 1970s w13x. Note that difficulties may arise for liquids with relatively large liquid surface tensions l , since the denominator of Eq. 55. can become zero. This limitation of the equation of state formulation is due to the use of Goods interaction parameter and is purely mathematical. Physical reasoning has been used to mend Eq. 55., and in practice Eq. 55. is implemented as a computer program w13x Appendix C. or a set of tables w127x. However, as mentioned, the use of Goods interaction parameter is not the only way to find an explicit expression for the equation of state; two different formulations have been discussed in the previous sections. It is instructive to compare Eq. 55. with that proposed by Li et al. w14,64x, Eq. 44., and the new formulation w125,126x, Eq. 50., by calculating s , with hypothetical values of l and . Examples of these results are given in Table 7. It is evident

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

221

that the three equations yield essentially the same s values. We wish to point out that these results are not trivial. The above three formulations were based on entirely different sets of experimental contact angle data: Eq. 55. was obtained using Zismans contact angles on eight polymer surfaces w13x; the in Eq. 44. was calibrated by the contact angles of Li et al. w14,64x on three different solid surfaces; the 1 in Eq. 50. was obtained from the low-rate dynamic contact angles on the 14 solid surfaces presented in Sections 2.3 and 2.4 w125,126x. It is remarkable that, despite the different formulations and experimental contact angles, Eqs. 44., 50. and 55. yield virtually the same s values, as given in Table 7. For practical purposes, any of the three equations can be used. 3.2.6. Predicti e power of the equation of state approach In this section, the predictive power of the equation of state approach will be illustrated simply by predicting, from a single contact angle measurement, all the other contact angles measured on one and the same solid surface. Since the s values obtained from the three formulations are virtually the same, the choice of any specific formulations becomes irrelevant. For example, if the surface tension l s 23.9 mJrm2 . and contact angle s 67.4 . of decane on a FC-722-coated mica surface are chosen from Fig. 7., the equation of state approach w14,64x Eq. 44., yields s value of 11.9 mJrm2 . From this s value, Eq. 44. can generate a series of contact angles for different
Table 8 Comparison between the measured and predicted contact angles and liquid surface tensions on a FC-722-coated mica Liquid degree. Measured Predicteda from Eq. 44.
l

mJrm2 . Predicteda from Eq. 44.

Measured

Decane 1-Pentanol trans-Decalin Hexadecane 1-Decanol cis-Decalin Ethyl cinnamate Dibenzylamine Dimethyl sulfoxide 1-Bromonaphthalene Diethylene glycol Ethylene glycol Diiodomethane 2,2 -Thiodiethanol Formamide Glycerol Water
a

67.36 72.95 73.38 75.94 78.84 79.56 86.54 90.70 90.95 93.81 94.22 97.87 101.18 104.56 108.49 111.73 118.69

71.4 73.4 74.2 76.4 81.3 87.4 91.5 93.5 95.2 95.6 98.4 100.7 106.3 108.7 113.5 119.3

23.88 26.01 27.19 27.62 28.99 32.32 37.17 40.80 42.68 44.31 44.68 47.55 49.98 56.26 59.08 65.02 72.70

26.8 27.2 28.6 30.4 31.3 36.6 40.2 40.4 43.3 43.5 47.0 50.5 54.4 59.0 62.9 71.9

Based on a

value of 11.9 mJrm2 calculated from the contact angle of decane.

222

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

values of l . These procedures are employed to produce a set of hypothetical contact angles corresponding to the surface tensions of the other liquids on the FC-722 surface. The predicted contact angle results are given in Table 8. In all cases, the predicted and measured contact angles agree well, generally within the usual error limit given in the literature as "2 . Conversely, a series of hypothetical l , values for different experimental contact angle values can also be generated from a single decane contact angle measurement. The predicted l values are also shown in Table 8 for the FC-722-coated surface; it can be seen that the deviations between the predicted and measured values vary from "0.1 mJrm2 to "2.0 mJrm2 : While these deviations are somewhat outside the experimental error limits, there is no evidence of any systematic variation. Similar calculations can be performed for other solid surfaces, with similar results. From a single contact angle measurement on one and the same solid surface, the equation of state approach w14,64x is capable of predicting the contact angles of different liquids as well as the liquid surface tensions.

4. Evaluation of existing contact angle data In Section 2.3, we have shown that static contact angles measured by conventional goniometer techniques could be meaningless in the context of Youngs equation. By measuring dynamic contact angles at very slow motion of the three-phase contact line using an automated drop shape technique, we obtained a variety of contact angle patterns, which cannot be easily detected by conventional goniometer contact angle techniques and procedures. For example, sliprstick of the three-phase contact line was obtained, where the contact angle increases steadily by as much as 35 at constant three-phase radius. Although such patterns are very interesting, they have to be discarded if the aim is to deduce solid surface tensions: since l , s , and sl are material properties and are expected to be constant, Youngs equation implies a unique contact angle. Thus, the interpretation of contact angles for solid surface tensions from sliprstick of the three-phase contact line would be meaningless. Apart from the applicability of Youngs equation, several other widely made assumptions are often overlooked. In an example Fig. 11. shown in Section 2.3.3, the contact angle increases as the drop front advances with essentially constant l . While there is no reason to question the applicability of Youngs equation, sl is suspected to change: from Youngs equation, if l and s are constant, changes in must be a consequence of a change in sl . While such cases might be very interesting in a different context, they have to be discarded if our goal is to deduce solid surface tension from the simplest possible situations. The above findings are not readily attainable for static contact angle measurements using conventional goniometer techniques and procedures: goniometer data may contain a mixture of meaningful and meaningless angles for solid surface tensions. If one neglects the goniometer angles shown to be meaningless in the low-rate dynamic contact angle study by, e.g. axisymmetric drop shape analysis

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

223

Fig. 21. l cos vs. l generated from the equation of state approach, Eq. 44., from mJrm2 to 45.0 mJrm2 , in 1.0 mJrm2 increments.

s 16.0

profile ADSA-P., smooth curves emerge in plots of l cos vs. l . This pattern was used to deduce a functional relationship to be used in conjunction with Youngs equation for determining solid surface tensions. Nevertheless, there is an abundance of contact angle data in the literature which were obtained from static measurements, and are hence of unknown status. The question then arises as to whether one should neglect completely the literature contact angle data and repeat all measurements using a dynamic procedure, such as ADSA-P, or whether some of the data can be salvaged. This question will be explored in this section w128x, using Zismans contact angle data on 34 different types of solid surfaces. The rationale of evaluating literature contact angle data is as follows. After discarding all those measurements which violate the basic assumptions made in all contact angle approaches of current interest w11 17x, the low-rate dynamic contact angle measurements yielded contact angles which fell on smooth curves in plots of vs. l , for one and the same solid surface w76,80,88 95x. Assuming that l cos this would also have happened in the systems studied in the laboratory of Zisman w10x, we propose to discard those contact angle data which do not sensibly fall on a smooth curve. The equation of state approach for solid liquid interfacial tensions w14x, i.e. Eq. 44., described in Section 3.2.3 can be used to evaluate these data by generating a series of theoretical constant s , curves, in plots of l cos vs. s . Fig. 21 shows the theoretical curves generated by the equation of state approach, i.e. Eq. 44., from s s 16.0 mJrm2 to s s 45.0 mJrm2 in 1.0 mJrm2 increments; each curve represents a constant s value and hence a hypothetical solid surface. Thus, if goniometer contact angle data are plotted in this graph, it can be decided how well data for any one solid fall onto an individual curve.

224

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Fig. 22. l cos vs. l for Zismans data on 80:20 copolymer of tetrafluoroethylene and chlorotrifluoroethylene w129x, 50:50 copolymer of tetrafluoroethylene and ethylene w129x, and polyvinyl fluoride w130x, as well as theoretical constanry, curves from Fig. 21.

Fig. 23. l cos vs. l for Zismans data on 80:20 copolymer of tetrafluoroethylene and chlorotrifluoroethylene w129x, 50:50 copolymer of tetrafluoroethylene and ethylene w129x, and polyvinyl fluoride w130x. Contact angle data which are clearly outside a constant s curve by more than "1.0 mJrm2 are discarded solid symbols..

Fig. 22 shows the contact angle data of Zisman et al. for 80:20 copolymer of tetrafluoroethylene and chlorotrifluoroethylene w129x, 50:50 copolymer of tetrafluoroethylene and ethylene w129x, and polyvinyl fluoride w130x surfaces, as well as the theoretical curves from Fig. 21. In Fig. 22, the data points for the 50:50 tetrafluo-

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

225

roethylene and ethylene surface fall consistently close to the theoretical constant 2 s curve of 27.0 mJrm . However, considerable scatter is apparent for the other surfaces. As we have discussed in Section 2.4.1, vapor adsorption of liquids which could cause s to change from one liquid to the next by as low as 1 mJrm2 has been considered possible w102x. If "1.0 mJrm2 is selected as the allowable error, contact angle data which deviate away from a theoretical s curve by more than "1.0 mJrm2 can be disregarded. It should be noted that this criterion is not an overestimate: the widely-accepted goniometer accuracy of "2 could easily cause the data points in Fig. 22 to vary along the l cos axis by as much as "2.0 mJrm2 . Following the above procedures, it was found that seven of the 22 data for the 80:20 tetrafluoroethylene and chlorotrifluoroethylene surface are clearly outside the error band of "1.0 mJrm2 and, similarly, six out of 11 data for the polyvinyl fluoride surface; these data are rejected, as shown in Fig. 23 in solid symbols. In principle, the above procedures can be employed for all other solid surfaces. However, if a large amount of contact angle data and solid surfaces are available, these procedures can be tedious and subjective. For this reason, they were automated by constructing a least-square objective function w123,128x to calculate the deviation of all contact angle data away from a theoretical curve generated by a given s value. The best-fitted constant s curve can then be used to discriminate against data which are clearly above or below the curve within a band, e.g. "1.0 mJrm2 . A FORTRAN computer program which automates these procedures is given in Appendix D. The computer program given in Appendix D was applied to the goniometer data for the following 34 solid surfaces from Zisman et al.: perfluorocapric acid w131x, perfluorolauic acid w131x, C 8 F17 SO 2 NC 3 H 7 . CH 2 CH 2 OOC CH CH 2 polymer S. w132x, C 7 F15 CH 2 OOC CCH 3 . CH 2 polymer A. w132x, perfluorocaprylic acid w131x, perfluorocaproic acid w131x, perfluorovaleric acid w131x, perfluorobutyric acid w131x, 17-perfluoroheptyl.-heptadecanoic acid w133x, 17-perfluoropentyl.-heptadecanoic acid w133x, perfluorodecanoic acid w134x, polytetrafluoroethylene w135x, 17-perfluoropropyl.-heptadecanoic acid w133x, paraffin w108x, hexatriacontane w108x, 17-perfluoroethyl.-heptadecanoic acid w133x, 80:20 copolymer of tetrafluoroethylene and chlorotrifluoroethylene w129x, stearic acid w136x, n-octadecylamine w137x, -monohydroperfluoroundecanoic acid w138x, 60:40 copolymer of tetrafluoroethylene and chlorotrifluoroethylene w129x, 50:50 copolymer of tetrafluoroethylene and ethylene w129x, polytrifluoroethylene w130x, polychlorotrifluoroethylene w129x, polystyrene w139x, 2-ethylhexanoic acid w140x, 2-ethylhexylamine w140x, 2-ethylhexanol w140x, polyvinylidene fluoride w130x, polyethylene terephthalate w139x, polyvinylidene chloride w130x, polyvinyl fluoride w130x, nylon 6,6. w139x, and perchloropentadienoic acid w130x. Fig. 24 shows the contact angle data for several solid surfaces, together with the fitted s curves; the rejected angles are given in solid symbols. It can be seen that, disregarding the rejected angles, the experimental data points fall closely to a distinct theoretical s curve for each solid within "1.0 mJrm2 . The contact angles for those pairs falling to "1.0 mJrm2 on smooth theoretical s , curves are given in Table 9, together with the calculated s values

226

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Table 9 Static contact angles from Zisman et al. which fall within "1.0 mJrm2 on a fitted Solid surface Liquid
l

curve
s

mJrm2 . degree. Equation of state, Eq. 44. Perfluorocapric acida w131x Octane Decane Dodecane Tetradecane Hexadecane Perfluorolauic acid w131x Pentamer Octane Decane Di- n-amyl. ether Dodecane Tetradecane Hexadecane Polyethylene V-120 Di- n-decyl. ether Benzene 1,6-Hexamethylene glycol Bis-2-ethylhexanoate Pentaerythritol tetracaproate Polyethylene SS903 8-p-Tolylnonadecane Bis-2-ethylhexyl . sebacate n-Decylbenzene 9- - cis-0,3,3-Bicyclooctyl.-methyl. heptadecane 4-Cyclohexyleicosane 1-Cyclohexyl-2-cyclohexylmethyl.pentadecane 1- -Decalylhendecane 9-n-Dodecylperhydro-phenanthrene 1,7-Dicyclopentyl-4-3cyclopentylpropyl.-heptane 1,3-Dicyclopentylcyclopentane 1,1-Di- -decalyl.-hendecane Benzylphenyl hendecanoate Tetrachlorobiphenyl Methylene iodide Tetramer Pentamer Hexamer Nonamer Tridecane Tetradecane Hexadecane Ethylene glyol Formamide Water 21.8 23.9 25.4 26.7 27.6 18.1 21.8 23.9 24.9 25.4 26.7 27.6 27.8 28.4 28.9 30.2 30.4 30.4 30.7 31.1 31.2 31.2 31.6 32.7 32.7 34.2 34.6 34.6 35.1 37.7 44.2 50.8 17.6 18.1 18.5 19.2 25.9 26.7 27.6 47.7 58.2 71.9 64.5 69.5 73.0 75.0 77.0 60.0 65.0 70.0 68.0 74.0 78.0 78.0 78.0 79.0 75.0 77.0 80.0 81.0 83.0 79.0 83.0 84.0 83.0 84.0 84.0 86.0 86.0 85.0 88.0 92.0 96.0 103.0 48.5 51.5 54.0 56.5 69.5 71.5 73.5 102.5 110.5 119.0 11.5 11.3 11.1 11.2 11.1 10.3 11.3 11.2 12.2 10.9 10.4 10.8 10.9 10.9 12.3 12.3 11.5 11.2 10.7 12.1 10.9 10.6 11.1 11.2 11.2 11.2 11.3 11.7 10.9 10.6 11.6 11.3 12.2 12.0 11.8 11.7 12.4 12.2 12.1 10.4 10.8 11.8

C8 F17 SO2 NC3 H7 . CH2 CH2 OOC CH CH2 polymer S. a w132x

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249 Table 9 Continued. Solid surface Liquid
l s

227

mJrm2 .

degree.

Equation of state, Eq. 44. 12.6 12.3 12.5 12.8 12.8 12.2 12.2 11.2 12.0 12.5 12.3 12.5 12.3 11.9 12.8 12.6 13.0 12.6 13.3 13.2 13.3 13.0 13.5 13.5 13.5 13.4 13.3 12.8 13.2 13.1 12.8 12.6 12.6 13.8 12.5 14.2 13.6 12.5 12.7 13.2 14.4 14.3

C7 F15 CH2 OOC CCH3 . CH2 polymer A.a w132x

Pentamer Hexamer Nonamer Dodecane Tridecane Tetradecane Hexadecane Formamide Water Octane Decane Dodecane Tetradecane Hexadecane Octane Decane Tetradecane Hexadecane Decane Dodecane Tetradecane Hexadecane Decane Dodecane Tetradecane Heptane Octane Decane Undecane Dodecane Tridecane Tetradecane 2-Octanol 1-Octanol Bis-2-ethylhexyl . phthalate Dicyclohexyl t-Butylnaphthalene Tricresyl phosphate Tetrachlorobiphenyl Formamide Glycerol Water

18.1 18.5 19.2 25.4 25.9 26.7 27.6 58.2 71.9 21.8 23.9 25.4 26.7 27.6 21.8 23.9 26.7 27.6 23.9 25.4 26.7 27.6 23.9 25.4 26.7 20.3 21.8 23.9 24.7 25.4 25.9 26.7 26.7 27.8 31.3 32.8 33.7 40.9 44.2 58.2 63.4 72.8

48.5 51.5 53.0 67.0 68.0 71.5 73.0 109.5 118.5 60.0 65.8 68.0 71.0 74.0 59.0 64.5 68.8 71.8 70.5 67.5 65.5 70.5 61.0 64.5 67.0 52.0 57.0 64.0 64.0 66.0 68.0 70.0 70.0 68.0 78.0 75.0 78.0 90.0 93.0 105.0 107.0 115.0

Perfluorocaprylic acida w131x

Perfluorocaproic acida w131x

Perfluorovaleric acida w131x

Perfluorobutyric acid w131x

17-Perfluoroheptyl.heptadecanoic acid w133x

228

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Table 9 Continued. Solid surface Liquid


l s

mJrm2 .

degree.

Equation of state, Eq. 44. 17.1 15.2 15.3 16.0 15.1 15.6 17.1 16.3 16.3 16.0 15.5 16.0 15.6 17.4 15.2 17.2 16.1 15.7 15.7 16.0 17.3 16.5 15.6 16.8 16.9 17.4 17.5 18.0 17.5 17.4 17.3 17.3 19.1 17.6 17.3 17.6 17.3 17.8 18.2 18.4 18.9 17.3 18.1 18.3

17-Perfluoropentyl.heptadecanoic acid w133x

Perfluoroalkane FCD-330. Dicyclohexyl t-Butylnaphthlene Tricresyl phosphate Methylene iodide Formamide Water Perfluorokerosene Di n-butyl. ether Diisoamyl. ether HCF2 .4 CH2 OH Di n-amyl. ether Carbon tetrachloride CFCl2 CF2 CCl3 sec-Butylbenzene Propylbenzene Butylbenzene Tris-2-ethylhexyl . Tricarballylate 1,6-Hexamethylene glycol Bis-2-ethylhexanoate t-Butylnaphthalene Perchlorocyclopentadiene Butylphenylhendecanoate Tricresyl phosphate Pentamethylene glycol Methylene iodide Tetramer Pentamer Hexane Hexamer Nonamer Dodecamer Heptadecamer Heptane Polyethylsiloxane Di-2-ethylhexyl. adipate Di-2-ethylhexyl. sebacate Di-2-ethylhexyl. phthalate Carbon disulfide t-Butylnaphthalene Tricresyl phosphate Tetraachlorodiphenyl Methylene iodide Glycerol Water

20.2 32.8 33.7 40.9 50.8 58.2 72.8 18.1 22.8 23.0 24.5 24.9 26.8 27.8 28.7 29.0 29.2 29.6 30.2 33.7 37.5 38.0 40.9 43.4 50.8 17.6 18.1 18.4 18.5 19.2 19.6 19.9 20.3 23.3 30.2 31.1 31.2 31.4 33.7 40.9 44.2 50.8 63.4 72.8

33.0 72.0 73.0 81.0 94.0 100.0 110.0 26.0 47.0 49.0 55.0 54.0 60.0 56.0 65.0 59.0 63.0 65.0 66.0 71.0 73.0 76.0 82.0 82.0 90.0 8.0 15.0 12.0 19.0 26.0 29.0 30.0 21.0 43.0 61.0 62.0 63.0 62.0 65.0 75.0 78.0 88.0 100.0 108.0

Perfluorodecanoic acid w134x

Polytetrafluoroethylene w135x

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249 Table 9 Continued. Solid surface Liquid
l s

229

mJrm2 .

degree.

Equation of state, Eq. 44. 18.6 18.4 18.6 19.6 18.1 18.8 19.4 20.2 20.1 20.2 18.2 18.7 19.3 19.0 19.9 20.1 19.0 19.8 18.6 19.1 19.6 19.6 19.7 19.0 19.6 19.9 20.2 18.6 20.0 20.1 20.3 20.8 20.7 21.1 21.1 21.0 21.0 21.3 20.2 20.9 19.7 20.2 20.2

17-Perfluoropropyl.heptadecanoic acid w133x

Perfluoroalkane FCD-330. Bis-2-ethylhexyl. phthalate Dicyclohexyl Tricresyl phosphate Ethylene glycol Methylene iodide Water Std. fluorolube Formamide Glycerol Water Polymethylsiloxane Di n-amyl. ether Std. fluorolube Di n-heptyl. ether Hexadecane Di n-decyl. ether n-Heptylic acid Tri2-ethylhexyl . tricarballylate Pentaerythtitol tetraeaproate Trieresyl phosphate Formamide Glycerol Dicyclohexyl Tricresyl phosphate Ethylene glycol Methylene iodide Formamide Water

20.2 31.3 32.8 40.9 47.7 50.8 72.8 25.1 58.2 63.4 72.8 19.9 24.9 25.1 27.0 27.6 27.7 28.3 29.6 30.4 40.9 58.2 63.4 32.8 40.9 47.7 50.8 58.2 72.8 20.3 20.5 22.8 22.9 24.9 25.4 27.0 27.7 28.8 31.2 37.7 50.8 58.2 63.4

23.0 60.0 62.0 72.0 84.0 86.0 106.0 38.0 91.0 96.0 108.0 20.0 41.0 43.0 45.0 46.0 50.0 49.0 56.0 56.0 72.0 92.0 97.0 61.0 72.0 80.0 83.0 94.0 105.0 8.0 8.0 25.0 26.0 33.0 35.0 41.0 43.0 45.0 54.0 64.0 84.0 91.0 96.0

Paraffin w108x

Hexatriacontane w108x

17-Perfluoroethyl.heptadecanoic acid w133x

80:20 Tetrafluoroethylene Heptane and chlorotrifluoroethylene Di n-propyl. ether w129x Di n-butyl. ether Nonane Di n-amyl. ether Dodecane Di n-heptyl. ether Di n-octyl. ether Benzene Di2-ethylhexyl . phthalene Benzyl phenylundecanoate Methylene iodide Formamide Glycerol

230

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Table 9 Continued. Solid surface Liquid


l s

mJrm2 . Stearic acid w136x

degree.

Equation of state, Eq. 44. 23.0 22.4 22.5 22.4 23.0 22.4 22.5 22.3 22.7 23.2 22.2 22.4 22.7 22.2 22.8 23.1 21.9 22.3 22.6 23.1 22.7 21.9 23.4 23.5 23.2 22.5 23.6 21.9 24.9 24.1 24.6 24.1 24.9 25.1 25.2 26.2 25.0 26.3 25.8 25.2 26.7

Octanoic acid Nonanoic acid Tricresyl phosphate Tetrachlorobiphenyl Thiodiglycol Glycerol Di n-butyl. ether Nonane Decane Di n-amyl. ether Dodecane Tetradecane Di n-heptyl. ether Hexadecane Di n-octyl. ether Di n-decyl. ether Tri2-ethylhexyl . tricarballylate Hexylbenzene 1,6-Hexamethylene glycol di-2-ethylhexanoate Pentaerythritol tetracaproate Di2-ethylhexyl. sebacate Di2-ethylhexyl. phthalate Decylbenzene -Methylnaphthalene Benzylphenylundecanoate Tricresyl phosphate Glycerol Water

29.2 29.5 40.9 44.2 54.0 63.4 22.8 22.9 23.9 24.9 25.4 26.7 27.0 27.6 27.7 28.4 29.6 30.0 30.2 30.4 31.1 31.2 31.2 36.4 37.7 40.9 63.4 72.8 25.2 30.2 31.2 63.4 72.8 27.0 27.6 28.8 33.7 40.9 44.6 63.4 72.8

40.0 43.0 65.0 70.0 81.0 92.0 9.0 13.0 18.0 22.0 30.0 34.0 34.0 38.0 36.0 37.0 45.0 39.0 44.0 43.0 46.0 49.0 44.0 55.0 58.0 65.0 90.0 102.0 9.0 39.0 40.0 89.0 97.0 22.0 24.0 25.0 45.0 56.0 63.0 87.0 94.0

n-Octadecylamine w137x

-Monohydroperfluoro- n-Amyl amine undecanoic acid w138x 1,6-Hexamethylene glycol di-2-ethylhexanoate Di-2-ethylhexyl . phthalate Glycerol Water 60:40 Tetrafluoroethylene and chlorotrifluoroethylene w129x Di n-heptyl. ether Hexadecane Benzene t-Butylnaphthalene Tricresyl phosphate -Bromonaphthalene Glycerol Water

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249 Table 9 Continued. Solid surface Liquid
l s

231

mJrm2 .

degree.

Equation of state, Eq. 44. 26.4 26.7 27.0 26.6 27.1 26.8 26.9 27.0 26.7 27.1 27.0 26.5 26.4 27.4 29.0 28.5 29.0 28.0 28.1 28.0 29.5 28.1 29.2 29.2 29.2 28.6 28.5 29.2 29.3 29.7 29.8 28.3 30.0 29.3 29.4 29.6

50:50 Tetrafluoroethylene and ethylene w129x

Tetradecane Di n-heptyl. ether Hexadecane Di n-octyl. ether Benzene Di2-ethylhexyl. phthalate t-Butylnaphthalene Benzyl phenylundecanoate Tricresyl phosphate -Bromonaphthalene Methylene iodide Formamide Glycerol Water Bis-2-ethylhexyl . phthalate Benzyl phenylundecanoate Tricresyl phosphate Formamide Glycerol Water Methylene iodide Glycerol Water Formamide Glycerol Water Decyl acetate Bis- n-butyl.-pyrotartrate Bis-2-ethylhexyl .- methyladipate 1,6-Hexanediol bis-2ethylhexanoate . Bis-2-ethylhexyl. tetrahydrophthalate Polyethylene SS 906 Bis-2-ethylhexyl.-sebacate 1-Cyclohexyl-2-cyclohexylmethyl.-pentadecane 9-n-Dodecylperhydrophenanthrene 1,1-Bis- -decalyl.-hendecane

26.7 27.0 27.6 27.7 28.8 31.2 33.7 37.7 40.9 44.6 50.8 58.2 63.4 72.8 31.2 37.7 40.9 58.2 63.4 72.8 50.8 63.4 72.8 58.2 63.4 72.8 28.5 29.3 30.2 30.2 30.7 30.7 31.1 32.7 34.2 35.1

9.0 9.0 12.0 16.0 20.0 32.0 39.0 48.0 55.0 60.0 69.0 79.0 85.0 93.0 22.0 44.0 49.0 76.0 82.0 92.0 64.0 82.0 90.0 74.0 80.0 91.0 2.0 5.0 14.0 10.0 14.0 23.0 15.0 27.0 32.0 34.0

Polytrifluoroethylene w130x

Polychlorotrifluoroethylene w129x

Polystyrene w139x

2-Ethylhexanoic acid w140x

232

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Table 9 Continued. Solid surface Liquid


l s

mJrm2 . degree. Equation of state, Eq. 44. 2-Ethylhexylamine w140x -Bromonaphthalene Methylene iodide 44.6 50.8 53.0 63.0 37.0 42.0 45.0 26.0 42.0 82.0 61.0 81.0 61.0 80.0 17.0 28.0 49.0 80.0 50.0 60.0 70.0 34.0 56.0 66.0 30.1 29.9 33.5 32.0 33.5 34.1 34.7 34.2 36.2 34.9 36.2 35.5 36.1 36.4 36.6 35.5 41.9 40.6 41.7 43.1 42.8 44.2

2-Ethylhexanol w140x

Tricresyl phosphate 40.9 Bis-2-phenylethyl.- -methyl adipate 41.3 -Bromonaphthalene 44.6 Benzyl phenylundecanoate -Bromonaphthalene Water Formamide Water Formamide Water Benzyl phenylendecanoate Tricresyl phosphate Methylene iodide Water Formamide Glycerol Water 37.7 44.6 72.8 58.2 72.8 58.2 72.8 37.7 40.9 50.8 72.8 58.2 63.4 72.8 50.8 63.4 72.8

Polyvinylidene fluoride w130x

Polyethylene terephthalate w139x Polyvinylidene chloride w130x Polyvinyl fluoride w130x

Nylon 6,6. w139x

Perchloropentadienoic acid Methylene iodide w130x Glycerol Water


a

Contact angles were estimated from graphs.

from the equation of state approach w14,64x wEq. 44.x. It can be seen that the calculated s , values for these surfaces are indeed quite constant. The s values for the 34 solid surfaces obtained from the above least-square fit are summarized in Table 10; the number of contact angles available and being rejected are also shown. To illustrate the procedures for eliminating contact angles, we reproduced in Fig. 25 the data for the polyvinyl fluoride surface from Fig. 22. For the 11 data in Fig. 25, a best-fitted s value of 36.3 mJrm2 was obtained from Eq. 44.. Thus, the contact angle which deviates away from a constant s curve generated by this value within a band will be rejected. The first rejected data point is 1 in Fig. 25; a new fitted s value was then obtained from the remaining 10 data. The above

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

233

Table 10 Summary of the s mJrm2 . values from curve fitting of Eq. 44. for 34 solid surfaces, using contact angle data from Zisman et al. Solid surface No. of data Available Perfluorocapric acida w131x Perfluorolauic acid w131x C8 F17 SO2 NC3 H7 . CH2 CH2 OOC CH CH2 polymer S.a w132x C7 F15 CH2 OOC CCH3 . CH2 polymer A.a w132x Perfluorocaprylic acida w131x Perfluorocaproic acida w131x Perfluorovaleric acida w131x Perfluorobutyric acid w131x 17-Perfluoroheptyl.-heptadecanoic acid w133x 17-Perfluoropentyl.-heptadecanoic acid w133x Perfluorodecanoic acid w134x Polytetrafluoroethylene w135x 17-Perfluoropropyl.-heptadecanoic acid w133x Paraffin w108x Hexatriacontane w108x 17-Perfluoroethyl.-heptadecanoic acid w133x 80:20 Tetrafluoroethylene and chlorotrifluoroethylene w129x Stearic acid w136x n-Octadecylamine w137x -Monohydroperfluoroundecanoic acid w138x 60:40 Tetrafluoroethylene and chlorotrifluoroethylene w129x 50:50 Tetrafluoroethylene and ethylene w129x Polytrifluoroethylene w130x Polychlorotrifluoroethylene w129x Polystyrene w139x 2-Ethylhexanoic acid w140x 2-Ethylhexylamine w140x 2-Ethylhexanol w140x Polyvinylidene fluoride w130x Polyethylene terephthalate w139x Polyvinylidene chloride w130x Polyvinyl fluoride w130x Nylon 6,6. w139x Perchloropentadienoic acid w130x
a s

Rejected 0 18 9 10 0 1 0 0 13 10 70 37 10 16 23 13 8 12 16 19 9 0 9 6 4 8 10 11 10 5 4 7 4 0 11.2 11.2 11.4 11.9 12.3 12.7 13.2 13.5 13.7 16.2 16.4 18.0 19.1 19.3 19.6 19.7 20.2 22.6 22.6 24.6 26.1 26.9 28.1 28.9 28.9 29.3 30.0 33.1 34.3 35.2 35.7 35.7 41.5 43.7

5 45 19 19 5 5 4 3 30 17 88 56 17 20 35 19 22 18 38 24 17 14 15 9 7 18 13 13 13 7 6 11 7 3

Contact angles were estimated from graphs.

procedures were repeated until all remaining. data points fall to "1.0 mJrm2 on a smooth s curve, as summarized in Table 11. Details have been given in Appendix D. For several solid surfaces in Table 10, s values had been obtained from more accurate contact angles. It is, therefore, instructive to compare these values with those determined from the above fitting procedures using Zismans goniometer data. For example, s values of 28.3 mJrm2 and 29.9 " 0.5 mJrm2 in Table 4.

234

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

Fig. 24. l cos vs. l for Zismans data on a. perfluorolauic acid w131x, b. 17-perfluoropentyl.heptadecanoic acid w133x, c. polytetrafluoroethylene w135x, d. n-octadecylamine w137x, e. polytrifluoroethylene w130x, and f. nylon 6,6. w139x. Contact angle data are discarded solid symbols. using the program given in Appendix D. The curves are best-fits of Eq. 44. onto experimental data of l and .

had been reported for two different types of polystyrene-coated silicon wafer surfaces w92,141x; they were determined from recent low-rate dynamic contact angles. In view of the difference in polymer properties and molecular weights, a value of 28.9 mJrm2 for the polystyrene given in Table 10 is indeed plausible. Furthermore, s s 35.6 " 0.4 mJrm2 for polyethylene terephthalate reported elsewhere w64x is again in good agreement with the value of 35.2 mJrm2 in Table 10. These findings give confidence in our procedures in deducing solid surface tension values from goniometer contact angles, so that dynamic contact angle procedures by, e.g. ADSA-P do not need to be repeated for the solid surfaces studied by Zisman et al. Thus, contact angles can indeed be used for the

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

235

Table 11 Summary of the sequence in rejecting contact angles for the polyvinyl fluroide surface w130x using the program given in Appendix D Sequence Data remaining. 1 11 2 11 3 11 4 11 5 11 5 8,10,11 5,6,8,10,11 5,6,8,11 Data rejected Fitted s mJrm2 . 36.3 36.5 36.6 36.7 36.7 36.2 36.1 35.7

1 2 3 4 5 6 7 8

1 2 3 4 9 7 10

Fig. 25. l cos vs. l for the polyvinyl fluoride w130x surface, reproduced from Fig. 23. The sequence of rejecting data is summarized in Table 11.

determination of solid surface tensions and literature contact angles do not have to be discarded completely.

5. Summary This study illustrates clearly that obtaining meaningful contact angles for the determination of solid surface tensions depends heavily on how contact angles are measured and whether or not the widely made assumptions have been violated. Experimental procedures and criteria for measuring and interpreting meaningful contact angles have been developed. The discrepancy in the literature in terms of

236

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

the contact angle patterns have been explained. A large amount of meaningful contact angle data was also generated on various solid surfaces. Unique experimental curves of l cos vs. l were obtained. The widely used surface tension component approaches w11,15 17x were shown to contradict these experimental contact angle patterns. The new contact angle data established here provide an experimental foundation for future development of the equation of state approach for solid surface tensions; a new formulation of such an approach which allows the determination of solid surface tensions has been developed. Literature static contact angles do not have to be discarded completely and can be used for determining solid surface tensions, with caution.

6. Perspective A look at the likely future of research on contact angles and surface energetics will benefit from a view back, particularly at persistent obstacles to progress. There are three key attitudes which have hampered progress immensely: 1. Contact angles are simple quantities which can be readily measured and interpreted by anybody, requiring no particular skill, methodology, or knowledge. 2. Contact angle measurements show hysteresis. Hence, experimental contact angles are not equilibrium values, and therefore nothing useful can be learned from contact angle studies. 3. Contact angles contain readily accessible information about intermolecular forces. All of these propositions are false. With respect to the first proposition, the present study shows how misleading conventional contact angle measurements on sessile drops can be. The second proposition, while false, is the most difficult to refute. Operationally, the key question is not whether a contact angle is an equilibrium contact angle, but whether it can be used in conjunction with Youngs equation. For instance, contact angle hysteresis caused by a modest degree of surface heterogeneity does not make Youngs equation inapplicable. Furthermore, it is clear from the systems examined here that contact with a liquid will often affect the solid surface. This does not preclude the possibility that the advancing contact angle is the equilibrium angle; rather this equilibrium cannot be reached from the other side, i.e. under receding conditions, because the solid surface is now different. Furthermore, it is difficult to see how the smooth curves of, e.g. Figs. 2 and 7, and Figs. 13 16 could arise if the contact angle data do not contain surface energetic meaning. Finally, the widely held view expressed in the third proposition is demonstrably false, from curves like those in Figs. 2 and 7, and Figs. 13 16. For a given solid, the contact angle depends only on the liquid surface tension, not directly on the intermolecular forces which give rise to these surface tensions. Referring to the curve for FC-721 in Fig. 2, the fact that methanol falls on the same smooth curve as hexane and decane would imply that the polar component of surface tension of

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

237

methanol is zero, i.e. methanol would have to be classified as non-polar, which is clearly absurd. While the first and third propositions have been dealt with extensively in this study, the answer to the question of contact angle hysteresis in the second proposition would require further study.

7. Future development The present state of knowledge is well summarized in the equation of state approach outlined above. The approach can be used to good advantage in numerous applications, including interfacial phenomena of powders w25,27x. Nevertheless, there remain many unsolved questions relating to the measurement and interpretation of contact angles. 1. While vapor adsorption does not appear to play a significant role in the contact angle data presented here, it is quite possible that it could enter the picture for more hydrophilic surfaces, particularly if l f s or l - s , i.e. cases which were excluded in the present study. While adsorption may produce contact angle patterns different from those shown in Figs. 2 and 7, and Figs. 13 16, this would not necessarily imply that Youngs equation is not applicable or that contact angle approaches might not be feasible. The answer to these questions would require general criteria to distinguish the effect of vapor adsorption on the contact angles from all other effects. 2. Recently, self-assembled monolayers SAMs. have been widely used to produce surfaces of different chemical compositions and wettabilities. However, it is believed that penetration of liquid into the SAMs is inevitable. Given that two solids having the same solid surface tensions and that a specific liquid may penetrate into one solid but not the other, the contact angles on the two chemically identical surfaces can be different: one would be due to pure energetic effects, and the other to the effects of the changed energetics and liquid penetration. In the latter case, attempts w109x to interpret surface energetics from contact angle approaches naively, e.g. by means of the above equation of state approach, could be misleading cf. Fig. 19.. 3. It is well-known that roughness affects the experimentally observed contact angles. However, there are as yet no general criteria to quantify roughness and at what level of smoothness surface topography has no longer an effect on the contact angle. The answer to such questions will have to involve considerations of line tension w59x and the contortions of three-phase lines. 4. It is still an open question of whether or not in Eq. 44. and 1 in Eq. 50. are universal constants, i.e. independent of the solid surface. Such a question can be addressed only after an even larger body of accurate contact angle data on various solids has been generated. 5. While the sliprstick contact angle patterns, e.g. that shown in Fig. 12 and

238

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

similar patterns cannot be used to determine surface energetics, they are still very interesting and worth investigating. 6. Routine production of systems which are free of contact angle hysteresis might be considered an important goal. It is firmly believed that all that would be produced is further confirmation of the contact angle story presented here. It is expected that more benefits can be obtained from the much simpler systematic study of receding contact angles of available systems, using concepts and procedures described in Section 2.3.

Acknowledgements This research was supported by the Natural Science and Engineering Research Council of Canada Grants: No. A8278 and No. EQP173469., Ontario Graduate Scholarships D.Y.K.., and University of Toronto Open Fellowships D.Y.K...

Appendix A This FORTRAN program may be used to calculate the solid liquid sl and solid vapor s interfacial tensions by the equation of state Eq. 44. with s 0.0001247 m2rmJ. 2 , using Newtons method. The required input data are the liquid surface tension l , and the contact angle . 10 implicit realU 8 a h,o z. write6,U . Input gamma LV and the contact angle deg. read 5,U . glv,theta const s 3.141592654r180.0 a s 0.5d0U dcosthetaU const. q 1.0d0. b s 0.0001247d0U glv U glv ysa xsy e s dexpbU 1.0d0 y xU x.UU 2. u s x y aU e v s 1.0d0 q 4.0d0U aU bU xU 1.0d0 y xU x.U e y s x y urv if y.lt. 1.0d0. go to 30 y s aU dexpb. go to 20 if dabsy y x..lt. 0.00001d0. go to 40 go to 20 gsv s xU xU glv gsl s gsv y glv U dcosthetaU const. cosin s dcosthetaU const. write 6,U . gamma LY ,glv, contact angle ,theta

20

30 40

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

239

50

write6,U . gamma SV ,gsv, gamma SL ,gsl write6,U . write6,U . Do you want to calculate again y s 1, n s 0. read5,U . num if num.eq. 1. go to 10 stop end

Appendix B This FORTRAN program may be used to calculate the solid liquid sl and solid vapor s interfacial tensions by the equation of state Eq. 50. with 1 s 0.0001057 m2rmJ. 2 , using Newtons method. The required input data are the liquid surface tension l , and the contact angle . 10 implicit realU 8 a h,o z. write6,U . Input gamma LV and the contact angle deg. read 5,U . glv,theta alpha s 0.0001057d00 b s alphaU glv UU 2 const s 3.141592654r180.0 a s 0.5d00U dcosthetaU const. q 1.0d0. x s 1.0d00 x1 s x call fx1,a,g,b,outf. call fprimeg,x1,b,outfp. x2 s x1 y outfroutfp if dabsx2 y x1..lt. 0.00001d0 to 40 x1 s x2 go to 30 gsv s glv U x2U x2 gsl s gsv y glv U dcosthetaU const. write 6,U . gamma LV ,glv, contact angle ,theta write6,U . gamma SV ,gsv, gamma SL ,gsl write6,U . Do you want to calculate again y s 1,n s 0. read5,U . num if num .eq. 1. go to 10 stop end subroutine fx1,a,g,b,outf. implicit realU 8 a h,o z. x s x1UU 2 g s 1.0d0 y bU 1.0d0 y x.UU 2 outf s x1U g y a

30

40

50

240

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

return end subroutine fprimeg,x1,b,outfp. implicit realU 8 a h,o z. x s x1UU 2 gprime s 4U xU bU 1.0d0 y x. outfp s g q gprime return end

Appendix C This FORTRAN program may be used to calculate the solid liquid sl and solid vapor s interfacial tensions, by the equation of state Eq. 55.. The required input data are the liquid surface tension l and the contact angle . 1 ns2 write6,U . Input gamma lv and contact angle deg.. read5,U .glv,angle call eqsaangle,glv,gsv,gsl. write6,U . gamma y 1v,theta,gamma y sv,gamma y sl write6,U .glv,angle,gsv,gsl write6,U . write6, U . Do you want to try again, Y s 1 read5,U .nn if nn .eq. 1. goto 1 stop end subroutine eqsaangle,glv,gsv,gsl. real angle,glv,gsv,gsl thedg s angle gaml s glv thema s thedg if gaml.lt.50. go to 980 diff s 0.7U gamlr5.. y 9..UU 2.. q 3 delth s abs90.y thedg. call largerthema,gaml,gams,gamsl. go to 983 980 themar s themaU 3.14159265r180 if thedg.gt.90.. go to 981 thedg s 180 y thedg 981 therd s thedgU 3.14159265r180 b s 0.015 cay s sdrt2.0 q bU gamlU costherd..

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

241

982 983

984

985

986

cophi s sqrtgamlU 27.U b.U 1.0 q costherd..r2.0U cay UU 3.0... ang s sqrt1.0 y cophiU cophi..rcophi phi s atanang. thibm s phi q 3.14159265.r3. gam s 4.0U cay U costhibm..UU 2.r3.0U B. if thema.gt.90.. go to 982 gamsl s gam gams s gamsl q gamlU costhemar. go to 983 gams s gam gamsl s gams y gamlU costherd. gsv s gams gsl s gamsl return end UUUUUUUUUUU UUUUUUUUUUU subroutine larger subroutine largertheta,gaml,gamsv,gamsl. real x150.,y150.,sg150. b s 0.015 h s gamlr2 is1 xi. s 90.0 yi. s 170. thedg s xi. q yi..r2. therd s thedgU 3.14159265r180. cay s sqrt2.0 q bU gamlU costherd.. cophi s sqrtgamlU 27.U b.U 1.0 q costherd..r2.0U cav UU 3.0... ang s sqrt1.0 y cophiU cophi..rcophi phi s atanang. thibm s phi q 3.14159265.r3. sgi. s y2.0U gamlU sin2.U thibm.U 1.y4.rbU gaml.. y 2..r 1.q costherd....r9.U ang. q 4.U gamlU costhibm.UU 2..r3. z s abssgi. y h. if z.le.0.00001. go to 986 if i.gt.149. go to 986 if sgi..lt.h. go to 985 xi q 1. s thedg yi q 1. s yi. isiq1 go to 984 xi q 1. s xi. yi q 1. s thedg isiq1 go to 984 gam s 4.0U camU costhibm..UU 2.r3.U b.

242

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

987

988

thddl s 180.y thedg thetr s thetaU 3.14159265r180. if theta.lt.thddl. go to 987 if theta.gt.thedg. go to 988 gamsv s gam q hU costhetr. y costherd.. go to 989 theda s 180.y theta thedr s thedaU 3.14159265r180. cay s sqrt2.0 q bU gamlU costhedr.. cophi s sqrtgamlU 27.U b.U 1.0 q costhedr..r2.0U cay UU 3.0... ang s sqrt1.0-cophiU cophi..rcophi phi s atanang. thibm s phi q 3.14159265.r3. gamsv s 2.0U cay U costhibm..UU 2..r3.0U b. q gamlU costhetr. go to 989 cay s sqrt2.0 q bU gamlU costhetr.. cophi s sqrtgamlU 27.U b.U 1.0 q costhetr..r2.0U cay UU 3.0... return end

Appendix D The FORTRAN program attached may be used to calculate the solid vapor interfacial tensions s by fitting goniometer contact angle data to theoretical s curves generated by the equation of state Eq. 44. using s 0.0001247 m2rmJ. 2 . The required input data are the liquid surface tension l and liquid surface tension times cosine of the contact angle l cos . Outputs are contact angles which fall within "1.0 mJrm2 to a best-fitted theoretical s curve, the s values corresponding to each data point, and a best-fitted s value to all data. The program reads experimental pairs of l and l cos from a file and then determines the smallest and largest s values for all data from Eq. 44.. For example, in the case of the polyvinyl fluoride surface, the smallest and largest values of s were found to be 28.7 mJrm2 1 in Fig. 25. and 39.9 mJrm2 9 in Fig. 25., respectively. They are needed for reasons of procedures to determine the best-fitted s value: it is assumed that the operative s will be somewhere in between these values. A first guess of the s value is obtained by fitting the entire set of experimental data to Eq. 44.. This yields a best-fitted s value of 36.3 mJrm2 , for the 11 data in Fig. 25. As an initial screening, "10.0 mJrm2 was selected as a criterion. Thus, data which deviate from s s 36.3 " 10.0 mJrm2 will be discarded first. The criterion is then decreased in "0.1 mJrm2 increments, i.e. "9.9 mJrm2 , "9.8 mJrm2 , etc. For each increment, contact angle data will be checked and those which exceed the band will be eliminated and a new fitted s value will be obtained from the remaining data. This procedure is repeated until

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

243

the band decreases to "1.0 mJrm2 . From the above example, as the band decreases, the first discarded data point is 1 in Fig. 25 and a new fitted s value of 36.5 mJrm2 was obtained based on the remaining 10 data. This s value was then used to eliminate contact angles by repeating similar procedures. The sequence of discarding data for the polyvinyl fluoride surface is summarized in Table 11. It turns out that only 4 data fall within "1.0 mJrm2 on a final best-fitted smooth s curve of 35.7 mJrm2 . This value is considered as the s value for the polyvinyl fluoride surface. implicit realU 8a h,o z. dimension glv100., ctheta 100., x100., y100., t100. dimension diff100., ndiff100., sgs100. character U 80 name write 6,U . Input data file name for gLV and gLV costheta read 5,U . name write 6,U . Input number of data points, n s ? read 5,U . n open unit s 1, status s old, file s name. do 5i s 1,n input gamma LV and gamma LV cos theta read 1,U .glvi.,ti. convert gamma LV cos theta to cos theta cthetai. s ti.rglvi. xi. s glvi. yi. s dlogdsqrtxi..U 0.5d0U 1.0d0 q ctheta i... continue close unit s 1. check for the lowest value of gSV from data sgs1. s eqsx1.,ctheta1.. gs1 s sgs1. do 55 i s 2,n sgsi. s eqsxi.,cthetai.. if sgsi. .lt. gs1. then gs1 s sgsi. end if continue check for the largest value of gSV from data sgs1. s eqsx1.,ctheta1.. gs2 s sgs1. do 51 i s 2,n sgsi. s eqsxi.,cthetai.. if gs2 .lt. sgsi.. then gs2 s sgsi. end if continue

c c

5 c

55 c

51

244

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

c 3 c

10

11

14 12

33

18 19 20 21

c c 10

best fit based on data difmax s 10.0 call fitsvx,y,n,ctheta,gs1,gs2,gsv. eliminate bad data based on fitted gSV nd s 0 do 10 i s 1, n diffi. s dabseqsxi.,cthetai.. y gsv. if diffi. .gt. difmax. then nd s nd q 1 ndiffnd. s i end if continue ifdifmax .le. 1.05d0. goto 33 difmax s difmax y 0.1d0 do 12j s 1,nd do 11i s ndiffj., n xi. s xi q 1. yi. s yi q 1. cthetai. s ctheta i q 1. continue do 14 i s j q 1,nd ndiffi. s ndiffi. y 1 continue continue n s n y nd goto 3 write6,19. do 18 i s 1,n write6,2D.xi.,xi.U ctheta i.,eqsxi.,cthetai.. continue write 6,21.gsv format ,r,FITTED RESULTS:,rr,5x,GLV,7x,GLV COS THETAq 4x,GSV. format ,3x,f6.2,8x,f6.2,7x,f6.2. format ,r,THE FITTED GSV s ,3x,f6.2. stop end subroutine fitsvx,y,n,ctheta,gs1,gs2,gsv. implicit realU 8 a h,o z. dimension ctheta 100., x100., y100. dimension gs500., R500. x s gamma LV b is defined as the beta b s 0.0001247d0 js1

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

245

gsj. s gs1 15 do 30 i s 1, n p1 s yi. y dloggsj..U 0.5d0 q bU xi. y gsj..UU 2 p2 s y0.5d0rgsj. y 2.0d0U bU xi. y gsj..UU 2 p3 s p1U p2 30 p4 s p4 q p3 Rj. s p4 ifdabsRj.. .lt. 0.000001. go to 100 p4 s 0.00 ifj .ne. 1.0. go to 40 gsa s gsj. Ra s Rj. jsjq1 gsj. s gs2 go to 15 40 ifj .ne. 2.0. go to 50 gsb s gsj. Rb s Rj. ifRaU Rb .lt. 0.0. go to 45 go to 10 45 j s j q 1 gsj. s 0.5U gsa q gsb. go to 15 50 ifRj.U Ra .lt. 0.0. go to 60 gsa s gsj. Ra s Rj. jsjq1 gsj. s 0.5U gsa q gsb. go to 15 60 gsb s gsj. Rb s Rj. jsjq1 gsj. s 0.5U gsa q gsb. go to 15 100 gsv s gsj. return end function eqsglv,ctheta. implicit realU 8 a h,o z. 10 a s 0.5d0U ctheta q 1.0d0. b s 0.0001247d0U glv U glv ysa 20 x s y e s dexpbU 1.0d0 y xU x.UU 2. u s x y aU e

246

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

30 40

v s 1.0d0 q 4.0d0U aU bU xU 1.0d0 y xU x.U e y s x y urv if y .lt. 1.0d0. go to 30 y s aU dexpb. go to 20 if dabsy y x. .lt. 0.00001d0. go to 40 go to 20 gsv s xU xU glv eqs s gsv return end

References
w1x w2x w3x w4x w5x w6x w7x w8x w9x w10x w11x w12x w13x w14x w15x w16x w17x w18x w19x w20x w21x w22x w23x w24x w25x w26x B.V. Derjaguin, V.M. Muller, Y.P. Toporov, J. Colloid Interface Sci. 73 1980. 293. K.L. Johnson, K. Kendall, A.D. Roberts, Proc. R. Soc. London. A324 1971. 301. V.M. Muller, V.S. Yushchenko, B.V. Derjaguin, J. Colloid Interface Sci. 92 1983. 92. A. Fogden, L.R. White, J. Colloid Interface Sci. 138 1990. 414. R.M. Pashley, P.M. McGuiggan, R.G. Horn, B.W. Ninham, J. Colloid Interface Sci. 126 1988. 569. H.K. Christenson, J. Phys. Chem. 90 1986. 4. P.M. Claesson, C.E. Blom, P.C. Horn, B.W. Ninham, J. Colloid Interface Sci. 114 1986. 234. P.M. Pashley, P.M. McGuiggan, R.M. Pashley, Colloids Surf. 27 1987. 277. R.M. Pashley, P.M. McGuiggan, B.W. Ninham, D.F. Evans, Science 229 1985. 1088. W.A. Zisman, Contact angle, wettability and adhesion, in: Advances in Chemistry Series, vol. 43, American Chemical Society, Washington, DC, 1964. F.M. Fowkes, Ind. Eng. Chem. 12 1964. 40. O. Driedger, A.W. Neumann, P.J. Sell, Kolloid-Z. Z. Polym. 201 1965. 52. A.W. Neumann, R.J. Good, C.J. Hope, M. Sejpal, J. Colloid Interface Sci. 49 1974. 291. J.K. Spelt, D. Li, The equation of state approach to interfacial tensions, in: A.W Neumann, J.K. Spelt Eds.., Applied Surface Thermodynamics, Marcel Dekker Inc., New York, 1996, pp. 239 292. D.K. Owens, R.C. Wendt, J. Appl. Polym. Sci. 13 1969. 1741. C.J. van Oss, M.K. Chaudhury, R.J. Good, Chem. Revs. 88 1988. 927. R.J. Good, C.J. van Oss, The modern theory of contact angles and the hydrogen bond components of surface energies, in: M. Schrader, G. Loeb Eds.., Modern Approaches to Wettability: Theory and Applications, Plenum Press, New York, 1992, pp. 1 27. H.G. Bruil, Colloid Polym. Sci. 252 1974. 32. G.D. Cheever, J. Coat. Technol. 55 1983. 53. H.W. Kilau, Colloids Surf. 26 1983. 217. K. Grundke, T. Bogumil, T. Gietzelt, H.-J. Jacobasch, D.Y. Kwok, A.W. Neumann, Prog. Colloid Polym. Sci. 101 1996. 58. E.I. Vargha-Butler, T.K. Zubovits, D.R. Absolom, A.W. Neumann, Dispers. Sci. Technol. 6 1985. 357. E.I. Vargha-Butler, E. Moy, A.W. Neumann, Colloids Surf. 24 1987. 315. E.I. Vargha-Butler, T.K. Zubovits, D.R. Absolom, A.W. Neumann, Chem. Eng. Commun. 33 1985. 255. D. Li, A.W. Neumann, Wettability and surface tension of particles, in: A.W. Neumann, J.K. Spelt Eds.., Applied Surface Thermodynamics, Marcel Dekker Inc., New York, 1996, pp. 509 556. S.N. Omenyi, A.W. Neumann, J. Appl. Phys. 47 1976. 3956.

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249

247

w27x D. Li, A.W. Neumann, Behavior of particles at solidification fronts, in: A.W. Neumann, J.K. Spelt Eds.., Applied Surface Thermodynamics, Marcel Dekker Inc., New York, 1996, pp. 557 628. w28x A.E. Corte, J. Geophys. Res. 67 1962. 1085. w29x P. Hoekstra, R.D. Miller, J. Colloid Interface Sci. 25 1967. 166. w30x J. Cisse, G.F. Bolling, J. Crystal Growth 10 1971. 67. w31x J. Cisse, G.F. Bolling, J. Crystal Growth 11 1971. 25. w32x A.M. Zubko, V.G. Lobonov, V.V. Nikonova, Sov. Phys. Crystallogr. 18 1973. 239. w33x K.H. Chen, W.R. Wilcox, J. Crystal Growth 40 1977. 214. w34x D.W. Fuerstenau, M.C. Williams, Colloids Surf. 22 1987. 87. w35x D.W. Fuerstenau, M.C. Williams, Part. Characterization 4 1987. 7. w36x D.W. Fuerstenau, M.C. Williams, Int. J. Mineral Process 20 1987. 153. w37x D.W. Fuerstenau, M.C. Williams, K.S. Narayanan, J.L. Diao, R.H. Urbina, Energy Fuels 2 1988. 237. w38x D.W. Fuerstenau, J. Diao, J. Hanson, Energy Fuels 4 1990. 34. w39x S.J. Hemingway, J.R. Henderson, J.R. Rowlinson, Faraday Symp. Chem. Soc. 16 1981. 33. w40x R. Guermeur, F. Biquard, C. Jacolin, J. Chem. Phys. 82 1985. 2040. w41x B.S. Carey, L.E. Scriven, H.T. Davis, AIChE J. 26 1980. 705. w42x E. Moy, A.W. Neumann, Theoretical approaches for estimating solid liquid interfacial tensions, in: A.W. Neumann, J.K. Spelt Eds.., Applied Surface Thermodynamics, Marcel Dekker Inc., New York, 1996, pp. 333 378. w43x H.C. Hamaker, Physica 4 1937. 1058. w44x J.N. Israelachvili, Proc. R. Soc. London. A. 331 1972. 39. w45x A.E. van Giessen, D.J. Bukman, B. Widom, J. Colloid Interface Sci. 192 1997. 257. w46x T.M. Reed, J. Phys. Chem. 55 1955. 425. w47x B.E.F. Fender, G.D. Halsey, Jr., J. Chem. Phys. 36 1962. 1881. w48x D.E. Sullivan, J. Chem. Phys. 74 1981. 2604. w49x D.V. Matyushov, R. Schmid, J. Chem. Phys. 104 1996. 8627. w50x T. Young, Philos. Trans. R. Soc. London 95 1805. 65. w51x A.W. Neumann, Adv. Colloid Interface Sci. 4 1974. 105. w52x D. Li, A.W. Neumann, Thermodynamic status of contact angles, in: A.W. Neumann, J.K. Spelt Eds.., Applied Surface Thermodynamics, Marcel Dekker Inc., New York, 1996, pp. 109 168. w53x A. Marmur, Colloids Surf., A 116 1996. 25. w54x R.V. Sedev, J.G. Petrov, A.W. Neumann, J. Colloid Interface Sci. 180 1996. 36. w55x R.N. Wenzel, Ind. Eng. Chem. 28 1936. 988. w56x A.B.D. Cassie, Discuss. Faraday Soc. 3 1948. 11. w57x S. Baxer, A.B.D. Cassie, J. Text. Inst. 36T 1945. 67. w58x A.B.D. Cassie, S. Baxter, Trans. Faraday Soc. 40 1944. 546. w59x J. Gaydos, A.W. Neumann, Line tension in multiphase equilibrium systems, in: A.W. Neumann, J.K. Spelt Eds.., Applied Surface Thermodynamics, Marcel Dekker Inc., New York, 1996, pp. 169 238. w60x C.J. van Oss, L. Ju, M.K. Chaudhury, R.J. Good, J. Colloid Interface Sci. 128 1989. 313. w61x C.J. van Oss, R.J. Good, J. Macromol. Sci-Chem. A26 1989. 1183. w62x A.W. Neumann, R.J. Good, Experimental methods, in: R.J. Good, R.R. Stomberg Eds.., Surface and Colloid Surface, vol. 11, Plenum Press, New York, 1979, pp. 31 91. w63x J.K. Spelt, E.I. Vargha-Butler, Contact angle and liquid surface tension measurements: general procedures and techniques, in: A.W. Neumann, J.K. Spelt Eds.., Applied Surface Thermodynamics, Marcel Dekker Inc., New York, 1996, pp. 379 411. w64x D. Li, A.W. Neumann, J. Colloid Interface Sci. 148 1992. 190. w65x D. Li, M. Xie, A.W. Neumann, Colloid Polym. Sci. 271 1993. 573. w66x Y. Rotenberg, L. Boruvka, A.W. Neumann, J. Colloid Interface Sci. 93 1983. 169. w67x P. Cheng, D. Li, L. Boruvka, Y. Rotenberg, A.W. Neumann, Colloids Surf. 43 1990. 151. w68x P. Cheng, Ph.D. Thesis, University of Toronto, 1990. w69x S. Lahooti, O.I. del Rio, P. Cheng, A.W. Neumann, Axisymmetric drop shape analysis, in: A.W.

248

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249 Neumann, J.K. Spelt Eds.., Applied Surface Thermodynamics, Marcel Dekker Inc., New York, 1996, pp. 441 507. D.Y. Kwok, D. Li, A.W. Neumann, Colloids Surf., A 89 1994. 181. C.A. Ward, A.W. Neumann, J. Colloid Interface Sci. 49 1974. 286. R. Defay, Etude Thermodynamique de la Tension Superficielle, Gauthier Villars, Paris, 1934. R. Defay, I. Prigogine, Surface tension and adsorption, A. Bellemans collab.., D.H. Everett trans.., Longmans, Green and Co., London, 1954, p. 222. D. Li, J. Gaydos, A.W. Neumann, Langmuir 5 1989. 293. D. Li, A.W. Neumann, Adv. Colloids Interface Sci. 49 1994. 147. D.Y. Kwok, T. Gietzelt, K. Grundke, H.-J. Jacobasch, A.W. Neumann, Langmuir 13 1997. 2880. D.H. Kaelble, J. Adhesion 2 1970. 66. S. Wu, in: L.-H. Lee Ed.., Adhesion and Adsorption of Polymers, Plenum, New York, 1980, pp. 53 65. C.J. van Oss, R.J. Good, M.K. Chaudhury, Langmuir 4 1988. 884. D.Y. Kwok, R. Lin, M. Mui, A.W. Neumann, Colloids Surf., A 116 1996. 63. J.F. Oliver, C. Huh, S.G. Mason, J. Colloid Interface Sci. 93 1983. 169. J.F. Oliver, C. Huh, S.G. Mason, Colloids Surf. 1 1980. 79. D. Duncan, D. Li, J. Gaydos, A.W. Neumann, J. Colloid Interface Sci. 169 1995. 256. A. Amirfazli, D.Y. Kwok, J. Gaydos, A.W. Neumann, J. Colloid Interface Sci. 205 1998. 1. D.Y. Kwok, D. Li, A.W. Neumann, Capillary rise at a vertical plate as a contact angle technique, in: A.W. Neumann, J.K. Spelt Eds.., Applied Surface Thermodynamics, Marcel Dekker Inc., New York, 1996, pp. 413 440. D.Y. Kwok, C.J. Budziak, A.W. Neumann, J. Colloid Interface Sci. 173 1995. 143. K. Grundke, P. Weidenhammer, P. Werner, R. Trinowitz, A. Janke, D.Y. Kwok, K. Poschel, H.-J. Jacobasch, A.W. Neumann, in preparation. D.Y. Kwok, C.N.C. Lam, A. Li, A. Leung, A.W. Neumann, Langmuir 14 1998. 2221. O.I. del Rio, D.Y. Kwok, R. Wu, J.M. Alvarez, A.W. Neumann, Contact angle measurements by axisymmetric drop shape analysis and an automated polynomial fit program, Colloid Surf. A 1998. in press. D.Y. Kwok, C.N.C. Lam, A. Li, A. Leung, R. Wu, E. Mok, A.W. Neumann, Measuring and interpreting contact angles: a complex issue, Colloids Surf. A 1998. in press. D.Y. Kwok, A. Leung, A. Li, C.N.C. Lam, R. Wu, A.W. Neumann, Colloid Polym. Sci. 276 1998. 459. D.Y. Kwok, C.N.C. Lam, A. Li, K. Zhu, R. Wu, A.W. Neumann, Polym. Eng. Sci. 38 1998. 1675. D.Y. Kwok, A. Li, C.N.C. Lain, R. Wu, S. Zschoche, K. Poschel, T. Gietzelt, K. Grundke, H.-J. Jacobasch, A.W. Neumann, Low-rate dynamic contact angles on polystyrene-alt-hexylr10carboxydecyl90r10.maleimide.. and the determination of solid surface tensions. In press. D.Y. Kwok, C.N.C. Lam, A. Li, A.W. Neumann, J. Adhes. 68 1998. 229. D.Y. Kwok, A. Leung, C.N.C. Lam, A. Li, R. Wu, A.W. Neumann, J. Colloid Interface Sci. 206 1998. 44. C.J. Budziak, A.W. Neumann, Colloids Surf. 43 1990. 279. G.H.E. Hellwig, A.W. Neumann, 5th International Congress on Surface Activity, Section B, 1968. 687. G.H.E. Hellwig, A.W. Neumann, Kolloid-Z. Z. Polym. 40 1969. 229. H.B. Callen, Thermodynamics and an Introduction to Thermostatistics, 2nd ed., John Wiley and Sons, 1985. G. Antonow, J. Chim. Phys. 5 1907. 372. D. Berthelot, Compt. Rend. 126 1898. 1857. R.J. Good, in: K.L. Mittal Ed.., Adsorption at interfaces, ACS Symposium Series, No. 8, American Chemical Society, D.C., 1975. D. Li, C. Ng, A.W. Neumann, J. Adhes. Sci. Technol. 6 1992. 601. C.D. Bain, E.B. Troughton, Y. Tao, J. Eval, G.M. Whitesides, R.G. Nuzzo, J. Am. Chem. Soc. 111 1989. 321. C.D. Bain, G.M. Whitesides, Angew. Chem. Int. Ed. Engl. 28 1989. 506.

w70x w71x w72x w73x w74x w75x w76x w77x w78x w79x w80x w81x w82x w83x w84x w85x w86x w87x w88x w89x w90x w91x w92x w93x w94x w95x w96x w97x w98x w99x w100x w101x w102x w103x w104x w105x

D.Y. Kwok, A.W. Neumann r Ad . Colloid Interface Sci. 81 (1999) 167 249 w106x w107x w108x w109x w110x w111x w112x w113x w114x w115x w116x w117x w118x w119x w120x w121x w122x w123x w124x w125x w126x w127x w128x w129x w130x w131x w132x w133x w134x w135x w136x w137x w138x w139x w140x w141x

249

P.E. Laibinis, G.M. Whitesides, J. Am. Chem. Soc. 113 1991. 7152. P.E. Laibinis, Ph.D. Thesis, Harvard University, 1991. H.W. Fox, W.A. Zisman, J. Colloid Sci. 7 1952. 428. J. Drelich, J.D. Miller, J. Colloid Interface Sci. 167 1994. 217. G.S. Ferguson, G.M. Whitesides, Thermal reconstruction of the functionalized interface of polyethylene carboxylic acid and its derivatives, in: M. Schrader, G. Loeb Eds.., Modern Approaches to Wettability: Theory and Applications, Plenum Press, New York, 1992, pp. 143 177. D.Y. Kwok, D. Li, A.W. Neumann, Langmuir 10 1994. 1323. W. Wu, R.F. Giese, Jr., C.J. van Oss, Langmuir 11 1995. 379. F. London, Z. Phys. Chem. B. 11 1930. 222. G.C. Maitland, M. Rigby, E.B. Smith, W.A. Wakeham, Intermolecular Forces: Their Origin and Determination, Clarendon Press, Oxford, 1981. A. Dupre, Theorie mecanique de la chaleur, Gauthier-Villars, Paris, 1969. L.A. Girifalco, R.J. Good, J. Phys. Chem. 61 1957. 904. R.J. Good, E. Elbing, Ind. Eng. Chem. 62 1970. 72. K.C. Chao, R.L. Robinson, Jr., Equation of State: Theories and Applications, ACS, Washington, D.C., 1986. J.S. Rowlinson, F.L. Swinton, Liquids and Liquid Mixtures, Butterworth Scientific, London, 1981. J. Kestin, E.A. Mason, AIP Conf. Proc. 11 1973. 137. D. Li, A.W. Neumann, J. Colloid Interface Sci. 137 1990. 304. E.W. Swokowski, E.W. Calculus with Analytic Geometry, 2nd ed., Prindle, Weber and Schmidt, Boston, 1979. C. Lipson, N.J. Sheth, Statistical Design and Analysis of Engineering Experiments, McGraw-Hill Publishing Co., New York, 1973. J.A. Mann, Jr., in: E. Matijjevic, R.J. Good Eds.., Surface and Colloid Science, vol. 13, Plenum Press, New York, 1984, p. 213. D.Y. Kwok, A.W. Neumann, Contact angle interpretation in terms of surface energetics, Colloids Surf. A, submitted for publication. D.Y. Kwok, Ph.D. Thesis, University of Toronto, 1998. A.W. Neumann, D.R. Absolom, D.W. Francis, C.J. van Oss, Sep. Purification Methods 9 1980. 69. D.Y. Kwok, A.W. Neumann, Re-evaluation of existing contact angle data in terms of surface energetics, Colloids Surf. A, submitted for publication. H.W. Fox, W.A. Zisman, J. Phys. Chem. 7 1952. 109. A.H. Ellison, W.A. Zisman, J. Phys. Chem. 58 1954. 260. E.F. Hare, E.G. Shafrin, W.A. Zisman, J. Phys. Chem. 58 1954. 236. M.K. Bernett, W.A. Zisman, J. Phys. Chem. 66 1962. 1207. E.G. Shafrin, W.A. Zisman, J. Phys. Chem. 66 1962. 740. F. Schulman, W.A. Zisman, J. Colloid Sci. 7 1952. 465. H.W. Fox, W.A. Zisman, J. Colloid Sci. 5 1950. 514. E.G. Shafrin, W.A. Zisman, J. Phys. Chem. 61 1957. 1046. E.G. Shafrin, W.A. Zisman, J. Colloid Sci. 7 1952. 166. A.H. Ellison, H.W. Fox, W.A. Zisman, J. Phys. Chem. 57 1953. 622. A.H. Ellison, W.A. Zisman, J. Phys. Chem. 58 1954. 503. H.F. Hare, W.A. Zisman, J. Phys. Chem. 59 1955. 335. A. Augsburg, K. Grundke, K. Poschel, H.-J. Jacobasch, A.W. Neumann, Determination of contact angles and solid surface tensions of poly4-x-styrene. films, Acta Polymerica, 1998. in press.

You might also like